21
Correlated variation and population differentiation in satellite DNA abundance among lines of Drosophila melanogaster Kevin H.-C. Wei, Jennifer K. Grenier, Daniel A. Barbash, and Andrew G. Clark 1 Department of Molecular Biology and Genetics, Cornell University, Ithaca, NY 14853-2703 Contributed by Andrew G. Clark, November 18, 2014 (sent for review October 1, 2014; reviewed by Giovanni Bosco and Keith A. Maggert) Tandemly repeating satellite DNA elements in heterochromatin occupy a substantial portion of many eukaryotic genomes. Al- though often characterized as genomic parasites deleterious to the host, they also can be crucial for essential processes such as chromosome segregation. Adding to their interest, satellite DNA elements evolve at high rates; among Drosophila, closely related species often differ drastically in both the types and abundances of satellite repeats. However, due to technical challenges, the evo- lutionary mechanisms driving this rapid turnover remain unclear. Here we characterize natural variation in simple-sequence repeats of 210 bp from inbred Drosophila melanogaster lines derived from multiple populations, using a method we developed called k-Seek that analyzes unassembled Illumina sequence reads. In ad- dition to quantifying all previously described satellite repeats, we identified many novel repeats of low to medium abundance. Many of the repeats show population differentiation, including two that are present in only some populations. Interestingly, the popula- tion structure inferred from overall satellite quantities does not recapitulate the expected population relationships based on the demographic history of D. melanogaster. We also find that some satellites of similar sequence composition are correlated across lines, revealing concerted evolution. Moreover, correlated satel- lites tend to be interspersed with each other, further suggesting that concerted change is partially driven by higher order structure. Surprisingly, we identified negative correlations among some sat- ellites, suggesting antagonistic interactions. Our study demon- strates that current genome assemblies vastly underestimate the complexity, abundance, and variation of highly repetitive satellite DNA and presents approaches to understand their rapid evolution- ary divergence. satellite DNA | population differentiation | rapid evolution H eterochromatin occupies a substantial portion of most eukaryotic genomes and contains vast quantities of tan- demly repeating, noncoding DNA elements known as satellite DNA. These sequences, along with transposable elements, are often described as selfish elements or genomic parasites, as they can increase their copy numbers irrespective of host fitness (1, 2). Indeed, they can be highly deleterious for the host genome; for example, ectopic recombination between homologous satellite repeats can lead to devastating chromosomal rearrangements (3, 4). Consequently, these elements are mostly sequestered in re- pressive chromatin environments around the centromeres and telomeres where there is minimal recombination and transcrip- tional activity. However, paradoxically, repetitive sequences are also crucial components of euchromatic genomes, as they recruit the centromeric histone H3 variant to form centromeres in many species (5, 6), thereby affecting the fidelity of chromosome seg- regation (7, 8). Adding to the perplexity, satellite DNA turns over at re- markably high rates between species (9, 10). In Drosophila mel- anogaster, satellite DNA is estimated to occupy over 20% of the genome. With the exception of the 359-bp (11), responder (12), and dodeca (13) satellites, most known satellites are tandem repeats of simple sequences (10 bp); the most abundant include AAGAG (aka GAGA-satellite), AACATAGAAT (aka 2L3L), and AATAT (11, 14, 15). In comparison, the genome of its sister species Drosophila simulans, from which D. melanogaster di- verged 2.5 mya, is estimated to have only 5% satellite DNA, more than 10-fold less AAGAG, and little to no AACATA- GAAT (16). For further contrast, nearly 50% of the Drosophila virilis and less than 0.5% of Drosophila erecta genomes are sat- ellite DNA (16, 17). Strikingly, such rapid changes in genomes have been implicated in postzygotic isolation of species in the form of hybrid incompatibility in several species of flies (1820), demonstrating the critical role satellite DNA has on the evolu- tion of genomes and species. The expansions and contractions of satellite sequences are thought to result from a combination of molecular events such as unequal crossing over (21), rolling circle replication (22), and polymerase slippage (23). Early population genetic studies as- sumed that satellite DNA has no function and that small changes in copy number are neutral, although total abundance may be under constraint due to the potential burden on metabolism, nuclear volume, and DNA replication (24). Under such assump- tions, early simulation studies demonstrated that unequal crossing over, drift, and reduced recombination are sufficient to generate long stretches of satellite DNA from random sequences (21, 25). Nevertheless, selection also appears to play an important role in shaping satellite DNA. For example, Stephan and Cho (1993) showed that selection is important in determining the length and heterogeneity of satellites, suggesting that the drastic interspecific Significance Most eukaryotic genomes harbor large amounts of highly re- petitive satellite DNA primarily in centromeric regions. Closely related Drosophila species have nearly complete turnover of the types and quantities of simple sequence repeats. How- ever, the detailed dynamics of turnover remains unclear, in part due to technical challenges in examining these highly repetitive sequences. We present a method (k-Seek) that identifies and quantifies simple sequence repeats from whole genome sequences. By characterizing natural variation in tandem repeats within Drosophila melanogaster, we identi- fied many novel repeats and found that geographically iso- lated populations show differentiation patterns that are, unexpectedly, incongruous with demographic history. More- over, repeats undergo correlated change in abundance, pro- viding additional insight into the dynamics of satellite DNA and genome evolution. Author contributions: K.H.-C.W., D.A.B., and A.G.C. designed research; K.H.-C.W. per- formed research; K.H.-C.W., J.K.G., D.A.B., and A.G.C. contributed new reagents/analytic tools; K.H.-C.W. analyzed data; and K.H.-C.W., D.A.B., and A.G.C. wrote the paper. Reviewers: G.B., Dartmouth College; and K.A.M., Texas A&M University. The authors declare no conflict of interest. 1 To whom correspondence should be addressed. Email: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1421951112/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1421951112 PNAS | December 30, 2014 | vol. 111 | no. 52 | 1879318798 POPULATION BIOLOGY

Correlated variation and population ... - Virginia Tech

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Correlated variation and population ... - Virginia Tech

Correlated variation and population differentiationin satellite DNA abundance among lines ofDrosophila melanogasterKevin H.-C. Wei, Jennifer K. Grenier, Daniel A. Barbash, and Andrew G. Clark1

Department of Molecular Biology and Genetics, Cornell University, Ithaca, NY 14853-2703

Contributed by Andrew G. Clark, November 18, 2014 (sent for review October 1, 2014; reviewed by Giovanni Bosco and Keith A. Maggert)

Tandemly repeating satellite DNA elements in heterochromatinoccupy a substantial portion of many eukaryotic genomes. Al-though often characterized as genomic parasites deleterious tothe host, they also can be crucial for essential processes such aschromosome segregation. Adding to their interest, satellite DNAelements evolve at high rates; among Drosophila, closely relatedspecies often differ drastically in both the types and abundancesof satellite repeats. However, due to technical challenges, the evo-lutionary mechanisms driving this rapid turnover remain unclear.Here we characterize natural variation in simple-sequence repeatsof 2–10 bp from inbred Drosophila melanogaster lines derivedfrom multiple populations, using a method we developed calledk-Seek that analyzes unassembled Illumina sequence reads. In ad-dition to quantifying all previously described satellite repeats, weidentified many novel repeats of low to medium abundance. Manyof the repeats show population differentiation, including two thatare present in only some populations. Interestingly, the popula-tion structure inferred from overall satellite quantities does notrecapitulate the expected population relationships based on thedemographic history of D. melanogaster. We also find that somesatellites of similar sequence composition are correlated acrosslines, revealing concerted evolution. Moreover, correlated satel-lites tend to be interspersed with each other, further suggestingthat concerted change is partially driven by higher order structure.Surprisingly, we identified negative correlations among some sat-ellites, suggesting antagonistic interactions. Our study demon-strates that current genome assemblies vastly underestimate thecomplexity, abundance, and variation of highly repetitive satelliteDNA and presents approaches to understand their rapid evolution-ary divergence.

satellite DNA | population differentiation | rapid evolution

Heterochromatin occupies a substantial portion of mosteukaryotic genomes and contains vast quantities of tan-

demly repeating, noncoding DNA elements known as satelliteDNA. These sequences, along with transposable elements, areoften described as selfish elements or genomic parasites, as theycan increase their copy numbers irrespective of host fitness (1, 2).Indeed, they can be highly deleterious for the host genome; forexample, ectopic recombination between homologous satelliterepeats can lead to devastating chromosomal rearrangements (3,4). Consequently, these elements are mostly sequestered in re-pressive chromatin environments around the centromeres andtelomeres where there is minimal recombination and transcrip-tional activity. However, paradoxically, repetitive sequences arealso crucial components of euchromatic genomes, as they recruitthe centromeric histone H3 variant to form centromeres in manyspecies (5, 6), thereby affecting the fidelity of chromosome seg-regation (7, 8).Adding to the perplexity, satellite DNA turns over at re-

markably high rates between species (9, 10). In Drosophila mel-anogaster, satellite DNA is estimated to occupy over 20% of thegenome. With the exception of the 359-bp (11), responder (12),and dodeca (13) satellites, most known satellites are tandem

repeats of simple sequences (≤10 bp); the most abundant includeAAGAG (aka GAGA-satellite), AACATAGAAT (aka 2L3L),and AATAT (11, 14, 15). In comparison, the genome of its sisterspecies Drosophila simulans, from which D. melanogaster di-verged ∼2.5 mya, is estimated to have only 5% satellite DNA,more than 10-fold less AAGAG, and little to no AACATA-GAAT (16). For further contrast, nearly 50% of the Drosophilavirilis and less than 0.5% of Drosophila erecta genomes are sat-ellite DNA (16, 17). Strikingly, such rapid changes in genomeshave been implicated in postzygotic isolation of species in theform of hybrid incompatibility in several species of flies (18–20),demonstrating the critical role satellite DNA has on the evolu-tion of genomes and species.The expansions and contractions of satellite sequences are

thought to result from a combination of molecular events such asunequal crossing over (21), rolling circle replication (22), andpolymerase slippage (23). Early population genetic studies as-sumed that satellite DNA has no function and that small changesin copy number are neutral, although total abundance may beunder constraint due to the potential burden on metabolism,nuclear volume, and DNA replication (24). Under such assump-tions, early simulation studies demonstrated that unequal crossingover, drift, and reduced recombination are sufficient to generatelong stretches of satellite DNA from random sequences (21, 25).Nevertheless, selection also appears to play an important role inshaping satellite DNA. For example, Stephan and Cho (1993)showed that selection is important in determining the length andheterogeneity of satellites, suggesting that the drastic interspecific

Significance

Most eukaryotic genomes harbor large amounts of highly re-petitive satellite DNA primarily in centromeric regions. Closelyrelated Drosophila species have nearly complete turnover ofthe types and quantities of simple sequence repeats. How-ever, the detailed dynamics of turnover remains unclear, inpart due to technical challenges in examining these highlyrepetitive sequences. We present a method (k-Seek) thatidentifies and quantifies simple sequence repeats from wholegenome sequences. By characterizing natural variation intandem repeats within Drosophila melanogaster, we identi-fied many novel repeats and found that geographically iso-lated populations show differentiation patterns that are,unexpectedly, incongruous with demographic history. More-over, repeats undergo correlated change in abundance, pro-viding additional insight into the dynamics of satellite DNAand genome evolution.

Author contributions: K.H.-C.W., D.A.B., and A.G.C. designed research; K.H.-C.W. per-formed research; K.H.-C.W., J.K.G., D.A.B., and A.G.C. contributed new reagents/analytictools; K.H.-C.W. analyzed data; and K.H.-C.W., D.A.B., and A.G.C. wrote the paper.

Reviewers: G.B., Dartmouth College; and K.A.M., Texas A&M University.

The authors declare no conflict of interest.1To whom correspondence should be addressed. Email: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1421951112/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1421951112 PNAS | December 30, 2014 | vol. 111 | no. 52 | 18793–18798

POPU

LATION

BIOLO

GY

Page 2: Correlated variation and population ... - Virginia Tech

differences may not be neutral (26). Since then, multiple authorshave emphasized the importance of both genetic drift and naturalselection in the evolution of repetitive DNA (27, 28). Further-more, recent studies have shown that repetitive sequences canhave remarkable effects on the rest of the genome. For example,natural variation in the Y chromosome, which is nearly entirelyheterochromatic, can modulate differential gene expression andcause variable phenotypes including differences in immune re-sponse (29, 30). These results reveal that changes in repetitivesequences can have fitness consequences on which selection willact. Meiotic drive models have also been proposed, in whichcentromeric satellites that bias the rate of transmission in femalemeiosis will quickly fix in the population (8, 31), providing anadditional mechanism for rapid turnover of satellite DNA.However, technical challenges have hindered research on het-

erochromatin. Because heterochromatic regions do not recom-bine, common genetic manipulations are mostly ineffective.Repetitive sequences, particularly low complexity satellite DNA,present severe challenges for making sequence assemblies andunique alignments (32). A handful of techniques have been ap-plied to study heterochromatin. High-density cesium–chloridegradient centrifugation has been instrumental in identifying majorsatellite blocks of different buoyancy, but fails to isolate lessabundant repeats (11). Hybridization approaches can only labelknown repeats and are often difficult to quantify precisely. Morerecently, flow cytometry has been used to indirectly estimateheterochromatic content, but it cannot distinguish the differenttypes of satellites contributing to the observed total (33).To address these shortfalls, we developed a computational

method, named k-Seek, that exhaustively identifies and quanti-fies short tandemly repeating sequences (kmers) from wholegenome sequences. We applied this method to 84 inbredD. melanogaster lines derived from natural populations and char-acterized the natural variation in satellite DNA. This allowed us toanswer three questions: (i) What are the abundances of all simpletandem repeat sequences in D. melanogaster? (ii) How are theirquantities changing within species and populations? (iii) How dothey change with respect to each other?

ResultsIdentification and Quantification of Tandem Repeats. We developedand validated a software package (k-Seek) that identifies andquantifies tandem repeats of 2 to 10mers from short read-basedwhole genome sequences (Fig. 1A). In short, each raw read isfirst broken into small fragments of equal lengths. Identicalfragments are then clustered. Whereas complex sequences areexpected to yield clusters with very few members, short repetitivefragments will form a large cluster. Once the kmer is identified,the number of repeats from the read is then tallied based ona word-search procedure. To capture tandem counts, only kmersthat are either immediately preceded or followed by the samekmer are scored. Additionally, we exclude tandem repeats thatspan less than 50 bp to avoid microsatellites and to guard againstascertainment bias for small kmers (2–4mers), as they are easierto identify from short stretches of DNA than larger kmers.Counts are summed across all reads and divided by the averageread depth of the uniquely mapped autosomal genome, allowingus to estimate the abundance of every identified kmer in the ge-nome (for detailed description, see SI Appendix). Benchmarkingwith simulated reads reveals that k-Seek is highly accurate atidentifying tandem repeats from 100-bp reads, and the counts arerobust against point mutations and indels (Fig. 1B and SI Ap-pendix, Fig. S1).To determine the reproducibility of k-Seek across library

preparations, we applied it to three DNA libraries independentlygenerated from line ZW155. The kmer quantities are highlycorrelated (Pearson’s r ranges from 0.976 to 0.988; SI Appendix,Fig. S2). Furthermore, the variability is not influenced by theabundance of the repeat (Fig. 1C). Nevertheless, several kmershave an elevated degree of variation. To independently assessthe accuracy of k-Seek, we quantified the abundance of the

10mer AACATAGAAT by measuring the radioactivity of [32]P-labeled probes hybridized to blotted DNA from 27 lines. Wefound significant correlation between the methods (Fig. 1D;Pearson’s r = 0.559, P < 0.005). We also attempted to quantifythe 5mer AAGAG using the same approach. However, thisprobe was problematic, and we were unable to obtain consistentresults across replicates and experiments (SI Appendix, Fig. S3).

Identification of Known and Novel kmers. We applied k-Seek toa collection of 84 inbred D. melanogaster lines sampled fromBeijing, Ithaca, Netherlands, Tasmania, and Zimbabwe, knownas the Global Diversity lines. Although there are 73,001 possible2–10mers, we only identified 72 distinct kmers with a populationmedian abundance of more than 100 bp per 1× depth across alllines (SI Appendix, Table S1). This list includes all previouslyidentified kmers (Fig. 2A and SI Appendix, Dataset S1). Asexpected, AAGAG and AACATAGAAT have the largest quan-tities, as they are two of the most abundant satellites known inD. melanogaster. Curiously, we detected AATAT at substantiallylower abundance than expected. This is likely due to under-amplification of sequences depleted of CGs during the PCR stageof library preparation (34). The most abundant kmer lengths were5mers and 10mers, whereas only a single 4mer was found. Mostbut not all (58/72) follow the (RRN)m(RN)n formula, where Rrepresents a purine andN represents any nucleotide, thought to becanonical for D. melanogaster satellites (11).Among the 72 kmers, 50 were previously unknown. Most are 5

and 10mers ranging from a normalized mean abundance of 105bp (AAGAGCAGAG) to 66,564 bp (AAGAT). Eleven of the 16new 10mers contain AAGAG, suggesting that they originatedfrom a mutation in one copy of AAGAG, followed by amplifi-cation of it and its nonmutated neighbor. Additionally, we findfour 8mers and six 9mers, lengths that had not previously beenidentified. Notably, most of the 8–9mers (7/10) do not follow the(RRN)m(RN)n formula and may therefore represent a qualita-tively distinct group of satellite sequences.

5mer repeats non-repeat

1.00.90.8

0.7

perfect subs indels

Freq

. of c

orre

ct

iden

tific

atio

n

2mer 3mer 4mer 5mer 6mer 7mer 8mer 9mer 10mer

A

B

1.2

0.8

0.4

0.2

100 10000 100000

Coe

ffici

ent o

f var

iatio

n

Normalized mean (bp)

C

Log 1

0(R

elat

ive

prob

e si

gnal

)

-0.4 -0.2 0 0.2 0.4

0.2

0

-0.2

-0.4

Log10(Relative abundance)

r = 0.559p < 0.005

Group1Group2 Group3 k-finder

k-counter

D

Fig. 1. The k-Seek package identifies and quantifies tandem kmers. (A)k-finder identifies kmers de novo by fragmenting short reads and groupingthem. k-counter then quantifies the number of tandem occurrences. (B)k-Seek applied to simulated 100-bp reads containing tandem arrays of kmersof different lengths. Simulated tandem arrays contained either perfectrepeats, up to four substitutions (subs), or indels. Frequency of correct iden-tification is plotted. (C) Variability of kmers between three independent li-brary preparations is plotted against the kmer abundance. Some of the highlyvariable kmers are labeled. Dotted line depicts the line of best fit. (D) Dot blotwith DNA from 27 lines hybridized with a probe targeting AACATAAGAT.Signal intensity is plotted against abundance inferred by k-Seek, both relativeto a reference line, with regression line plotted in red. Error bars are SEscalculated from three replicates of each sample in the dot blot.

18794 | www.pnas.org/cgi/doi/10.1073/pnas.1421951112 Wei et al.

Page 3: Correlated variation and population ... - Virginia Tech

Population Structure. Across all Global Diversity lines, the aver-age total kmer count is 4.03 Mb per 1× read depth. Strikingly,the lowest and highest lines differ by 2.50-fold (equating to 4.29Mb difference), indicating high intraspecific variability (SI Ap-pendix, Fig. S4). AAGAG and AACATAGAAT, the two mostabundant kmers, comprise 74% of the total kmer counts onaverage but can be as high as 88% and as low as 57%, revealingmarked differences in the repeat composition among lines.The phylogenetic relationships of the lines were inferred from

genome-wide SNP calls, and that analysis largely recapitulatesthe expected demographic history of D. melanogaster, with anAfrican origin and a relatively recent global spreading alonghuman trade routes (for review, see ref. 35). The simple expec-tation is that kmer abundance will also reflect the same pop-ulation structure. However, hierarchical clustering of kmerabundances failed to differentiate the lines into their respectivepopulations (SI Appendix, Fig. S5). To further investigate, weapplied principal components analysis on the top 100 kmers (Fig.3A). The Zimbabwe lines fall into a diffuse cluster with minimaloverlap with other populations, as expected. The Netherlandslines broadly cluster with the Tasmanian lines, consistent withthe introduction of D. melanogaster to Australia by Europeansettlers (36). Surprisingly, although, the Beijing lines largelyoverlap with the Ithaca lines, even though North Americanpopulations are thought to be of European origin (36, 37) anddistinct from Asian populations established shortly after theinitial out-of-Africa migration (38). These discrepancies suggestthat satellite DNA abundance is subject to a distinct evolutionaryhistory from the rest of the genome.To infer the population differences for each kmer, we applied

Rst statistics, assuming a step-wise mutational model (Fig. 3B)(39, 40). Of the top 100 kmers, the majority (n = 54) display verylittle population differentiation; many (n = 46) have an Rst of>0.1, showing appreciable population differentiation; and some(n = 7) have an Rst of >0.4, revealing high population differ-ences. For example, the AT 2mer has a startlingly high Rst of0.554, which appears to be due to elevated levels in the Neth-erlands and Tasmania populations (Fig. 3C). AAGAG andAACATAGAAT, the most abundant kmers, have moderatelevels of differentiation, with the Zimbabwe population havingthe highest and lowest abundance, respectively (Fig. 3 D and E).Because different kmers have distinct patterns of populationdifferentiation, we conclude that they experience different evo-lutionary dynamics.Among the most differentiated kmers, two 10mers (AACA-

TATAAT and AAAATAGAAT) are surprisingly found only inthe Netherlands, Tasmania, and Zimbabwe populations, whilebeing completely absent in the Beijing and Ithaca populations(Fig. 3F). Within the Tasmania and Netherlands populations,there is high variation even among individual lines, which sug-gests high turnover rates. To confirm the presence/absence ofpolymorphism, we designed FISH probes targeting AACATA-TAAT and observed fluorescent foci from mitotic chromosomesof the Netherland and Tasmania but not the Beijing lines (Fig.3G and SI Appendix, Fig. S6). The foci are autosomal and appearto be centromeric, as they are located near the primary con-striction and do not overlap with the predominantly pericentricAAGAG foci. This finding provides, to our knowledge, the firstreport of population-specific satellite DNA in Drosophila andfurther underscores the high rate of satellite DNA turnover.

Concerted Evolution of kmer Abundance. Interestingly, the twopopulation-specific 10mers are highly positively correlated acrosslines (Fig. 3G; Pearson’s r = 0.993, P < 2.2 × 10−16), suggestingthat they undergo coordinated changes in copy number. Tocomprehensively identify kmers that are evolving in a concertedfashion, we generated a pairwise correlation matrix for the top100 kmers and clustered those that are highly correlated (Fig. 4Aand SI Appendix, Fig. S7). This was accomplished using Modu-lated Modularity Clustering, which rearranges rows and columnsof the correlation matrix to identify clusters of variables withmaximal pairwise correlations among all cluster members (in thiscase, kmers) without predetermined knowledge or an arbitrarydecision on number of clusters (41). Overall, we find nine majorclusters of correlated kmers. As expected, the two population-specific 10mers are found within the same cluster along with theAATAT 5mer (Fig. 4C). Clustering appears to be driven in partby sequence similarity; several clusters are either AT-rich, AG-rich,or AC-rich. For example, the AG-rich cluster contains AAGAG,as well as related sequences AAGAGAG and AAGAGAGAG.However, many highly related kmers fall into separate clusters—for example, AACATAGAAT (the most abundant 10mer) andAACATATAAT (one of the two population-specific kmers)—even though they only differ by one nucleotide. Surprisingly, wealso observe relatively weak but significant negative correlationsamong a small number of kmers (Fig. 4A and SI Appendix, Fig.S8). Notably, the AG-rich kmers are negatively correlated withthe AT-rich kmers; not only are the two respective clusters anti-correlated (Fig. 4A, arrowhead), the 10mer AAGAGCAGAGthat is grouped within the AT-rich cluster is also negatively cor-related with all other AT-rich kmers (Fig. 4B). These negative

100001000100101

0.10.01

0.001

Nor

mal

ized

Med

ian

(kb)

Fig. 2. kmer abundance. Medians of the top kmersacross all strains are plotted in log10 scale. Error barsrepresent the first and third quartiles. Gray and redbars are even- and odd-number kmers, respectively.Previously characterized kmers are labeled in bold.kmers not following (RRN)m(RN)n are labeled by anasterisk. AAGAG-containing 10mers are in red.

A B

C D

E F

G

Fig. 3. Population structure of kmers. (A) Lines are plotted based on the firstthree principal components derived from the top 100 kmers. Lines from the samepopulations are circled with the respective colors. (B) Distribution of populationdifferentiation index RST. (C–E) Distribution of abundance of selected kmers inthe five populations. (F) Abundance of AACATATAAT and AAAATAGAAT acrosslines. (G) FISH applied to mitotic chromosomes of lines from Beijing and Tas-mania. Probes for AAGAG are labeled green, AACATATAAT red, and DAPI blue.Arrowheads indicate red foci.

Wei et al. PNAS | December 30, 2014 | vol. 111 | no. 52 | 18795

POPU

LATION

BIOLO

GY

Page 4: Correlated variation and population ... - Virginia Tech

correlations suggest that kmers can have antagonistic relationshipssuch that expansion of one comes at the expense of another.

Interspersion of kmer Blocks Drives Correlation. One possible causeof the observed positive correlations is that correlated kmersrepresent physically linked and interspersed satellite blocks.Therefore, deletions or duplications of these repetitive blockswill decrease and increase both kmers in concert. To test thispossibility, we identified, across all lines, paired-end reads wherekmers are found in both mate pairs and determined the fre-quency of their occurrences relative to the abundance of theidentified kmers (Fig. 5A). As expected, almost every kmer ismost frequently paired with itself, reflecting that many comprisesizeable and homogenous blocks. However, kmers found withinpositively correlated clusters tend to be found in mate pairs morefrequently than those outside (Fig. 5 B–D), consistent with ourhypothesis. For example, kmers within the AG-rich cluster arehighly interspersed with one another. This is further supportedby a significant and positive association between the correlationvalues and the interspersion frequency of the kmers (Fig. 5E).The two population-specific 10mers identified are also highly

interspersed (Fig. 5C). Interestingly, they are most frequentlypaired with each other, and mate pairs containing the same10mer are rarely found. This result suggests that the two 10mersare interspersed with each other in small blocks that are roughlythe length of the insert size, which is ∼450 bp (Fig. 5A). This isalso true for the AAC 3mer and the AAAATAACAT 10mer,suggesting that they also exist in small interdigitated blocks.We note that there are many instances where interspersed

kmers are not correlated. This is unsurprising, as interspersionitself is insufficient to drive correlated change if the blocks do notexperience duplication and/or deletion. Additionally, for kmersthat are found interspersed with many other kmers, presumablyin separate blocks, independent indels in different blocks willresult in local concerted change in abundance, but their abun-dances aggregated across the genome will likely be uncorrelated.Of further interest are correlated kmers that are not interspersed,such as the two population-specific 10mers and the 5merAATAT, as they indicate additional mechanisms underlying theconcerted change. However, it is difficult to distinguish thesefrom interspersion that we fail to capture due to low coverage orunderrepresentation. Furthermore, any junction between satelliteblocks that is gapped by complex sequences, such as transposableelements, will also likely be missed.

DiscussionMany important questions in heterochromatin biology are nowaccessible using our software pipeline (k-Seek). Previously, iden-tification of satellite sequences was mostly accomplished throughlabor-intensive methods that have low sensitivity (16). As a result,the current catalog of satellite DNA contains exclusively kmersthat are present in large quantities. Our method is accurate atidentifying tandem kmers and discovered many previously un-known kmers of low to medium abundance. We expected thatPCR would be a major source of bias during library preparation asthe polymerase underamplifies AT-rich sequences, and we indeedfound lower abundance of AATAT compared with previouscharacterizations (11). Nevertheless, using three replicate librariesmade from a single sample, we found such bias to be consistentacross the libraries, allowing us to characterize population varia-tion of individual kmers.

Potential Causes of Population Variation. By applying k-Seek to theDrosophila Global Diversity lines, we characterized natural var-iation in heterochromatin repeat structure. The mean satelliteabundance in a population is expected to approximate an equi-librium determined by the mutation rate, the degree of selectiveconstraint, potentially positive selection, and population size(24, 25). For many kmers, the difference between populations issmall, and the low Rst suggests a high rate of migration orturnover. Nevertheless, we identified multiple kmers with ap-preciable to high population differentiation. Notably, the patternof interpopulation differentiation is also variable among repeats,revealing that some kmers evolve relatively independently ofothers. The process driving the population differences could beeither neutral drift or natural selection. According to the out-of-Africa model, the Zimbabwe population is expected to have thehighest level of genetic variation, provided that the differencesare nearly neutral. Although this is, as expected, true for allkmers considered together (SI Appendix, Fig. S9), we foundmany exceptions that may be revealing of their modes of evo-lution. For example, the Netherlands population not only hasa significantly higher abundance of the AT 2mer compared withZimbabwe, but the between-line variability is also substantially

0 1-1

A B

C

D

Fig. 4. Correlation structure of kmer variation. (A) Pair-wise correlationmatrix is reorganized such that kmers with correlated change across lines areclustered into groups demarcated by boxes. Colors represent strength ofSpearman’s correlation. Arrowhead indicates clusters that are negativelycorrelated. (B–D) Magnified clusters with AT-rich, the two population-spe-cific (bold), and AG-rich kmers, respectively. For the fully labeled matrix, seeSI Appendix, Fig. S6.

(Non-junction reads) Tandem blocks Interspersed small blocks

(Junction reads)

Interspersed blocks

Interspersed Freq. 1 1e-4 1e-8 1e-12

A B C

D

1e-1 1e-2 1e-3 1e-4 1e-5 1e-6

0.0 0.2 0.4 0.6 0.8

Inte

rspe

rsed

Fre

q.

Spearman’s ρ

E

Fig. 5. Interspersed kmer blocks. (A) Paired-end reads are used to inferinterspersion. Interspersed kmers will have many mate pairs spanning thejunctions (Left). Large kmer blocks are in tandem and will have few matepairs containing different kmers (Middle). Interspersion of small blocks(∼450 bp) will yield only mate-pair gapping junctions (Right). (B) Inter-spersion frequency matrix for kmers organized as in Fig. 4A. (C and D)Magnified cluster containing the two population-specific 10mers (in bold)and AG-rich cluster, respectively. (E ) The correlation strength for eachkmer pair (from Fig. 4A) is plotted against the interspersed frequency(from B). Self-correlation and interspersion (cells across the diagonal) andpairs with no interspersion are excluded. Regression line is plotted in red(Pearson’s r = 0.325, P = 9.77 x 10−15).

18796 | www.pnas.org/cgi/doi/10.1073/pnas.1421951112 Wei et al.

Page 5: Correlated variation and population ... - Virginia Tech

greater. These differences may be indicative of a relaxation ofconstraint within the Netherlands population, allowing for labileexpansion and contraction. In contrast, the differentiation pat-tern of AAGAG shows significant reduction in the non-Africanpopulations, potentially reflecting an increase in the level ofselective constraint after the out-of-Africa migration, or reducedvariability due to the out-of-Africa bottleneck as is seen for mostof the genome (42).The incongruity between the population structure inferred

from kmer abundance and from demographic history and SNPsis intriguing. This is reminiscent of the well-documented phe-nomenon of the homogenization of multicopy gene families andtandemly arrayed genes such as rDNA, a process that has beencalled “molecular drive” (43). Resulting from sequence exchangesvia gene conversion, paralogs that predate the species split maydisplay a high degree of within-species sequence homogeneity.Depending on the stochastic or biased dynamics of the process,phylogenetic relationships of arrays between species and by ex-tension populations may not be preserved in these sequences.Alternatively, the discrepancy may be due to incomplete lin-

eage sorting (44) of some repeats. In one possible scenario,individuals without the correlated AT-rich repeats segregated atlow frequency in the European population and were subse-quently introduced by chance only to North American but not toAustralia. Additionally, we cannot rule out the possibility thatthe similarity between the Beijing and Ithaca lines is due to se-lection from a common environmental pressure.Meiotic drive and segregation distortion present an additional

explanation of strong population-specific patterns. In these sce-narios, chromosomes with a particular kmer abundance or com-position can have a segregation advantage in some populations. Ifthese repetitive sequences have pleiotropic deleterious effects,fixation of a suppressor can quickly purge these sequences frompopulations. Our discovery of the two population-specific 10mers(Fig. 3G) is particularly striking and suggestive of the rapid evo-lution predicted by most models of meiotic drive. Notably, theselines provide genetic material for direct empirical tests of possiblesegregation differences (now underway).

Concerted Change of kmer Abundances. We identified severalgroups of highly correlated repeats, revealing that different sat-ellites undergo concerted evolution in abundance. The fact thatkmers within a correlated group have sequence similarity issuggestive of the potential underlying mechanism. Several DNAbinding proteins have been identified to target satellite DNA(for review, see ref. 45). GAGA-factor is a transcription factorresponsible for key developmental regulation (46), heat-shockresponse, and chromatin remodeling (47, 48) that localizes toAAGAG and AAGAGAG satellites during mitosis (49, 50).Here, we showed that these two are correlated along with otherAG-rich kmers, thus raising the possibility that binding toa common protein drives concerted change. In one scenario, anincrease in GAGA-factor more effectively packages AG-richrepeats into heterochromatin, thereby raising host tolerance tothe repeats. Subsequently, both repeats will increase in number.Conversely, a decrease in the protein level may result in sub-optimal regulation of repeats and reduction in organismal fit-ness; therefore, selection will favor individuals with less of thetargeted repeats, resulting in correlated contraction. Similarly,the concerted change between AT-rich kmers may reflect pro-teins that recognize AT-rich satellites, including origin recognitioncomplex subunit 2, an essential component of the complex thatinitiates DNA replication (51), and D1, an essential protein im-plicated in chromatin remodeling (52). Additionally, the PRODprotein (proliferation disrupter) binds to the AACATAGAAT10mer on mitotic chromosomes (53). Our results suggest that itmay also bind to other satellites, as several kmers are correlatedwith this 10mer.However, the observed concerted changes are not necessarily

driven by selection. We demonstrate that the structure of kmerscan also account for some of the observed correlated patterns.

For satellite blocks that are interspersed with each other, du-plications or deletions of the region will increase or decrease thekmers together. Indeed, kmers that are highly interspersed tendto have higher correlation in abundance, and the high inter-spersion of the two population-specific 10mers serves as a primeexample. AAGAG and AAGAGAG are also moderately inter-spersed, suggesting that GAGA-factor binding may not be theonly mechanism driving their correlation. We therefore concludethat the complex architecture of satellite DNA is likely the resultof both neutral and selected mutational changes.Surprisingly, we also identified negative correlations, albeit

weak ones, notably between the AT- and AG-rich kmers. Wespeculate that such antagonistic interactions reflect an optimalload of satellite DNA that genomes can tolerate. Therefore, thedeleterious effect of an increase in one satellite DNA group canbe alleviated by a decrease in a different group. This load may bedetermined by the fitness benefits of maintaining an optimalgenome size (24), but we find this unlikely given the variability ofgenome sizes among as well as within species (54) and that totalkmer quantities differ greatly between lines. Alternatively, theload may be chromosome-specific, as satellite DNAs often havechromosome-specific distributions. An optimal load of satelliteDNA may ensure faithful chromosomal transmission or preventdeleterious rearrangements, as lengthy satellite blocks may bemore prone to unequal crossing over or ectopic recombination.Regardless of the specific molecular and evolutionary mecha-nism, the observed antagonistic relationships intimate the curi-ous possibility that satellite DNAs are at odds not only with thehost genome but also with each other.

Materials and MethodsFly Lines and Sequence Reads. The 84 lines ofD.melanogaster used in this studywere sib-mated from isofemales lines (55). Whole genome shotgun sequencingwas done on the Illumina platform with DNA extracted from adult females,sequenced to a depth of 12.5× for each line using paired-end 100-bp reads.

Normalizing Tandem kmer Counts. To normalize kmer counts between lines,we divided all counts by the average read depth at autosomal regions andthen multiplied the counts by the kmer length, to obtain number ofnucleotides per 1× depth (SI Appendix, Dataset S1). Average read depth wasobtained by mapping sequences with BWA to D. melanogaster referencer.546 (Flybase). We used Picard tools to compute the read depth distributionand averaged across the autosomes. We note that very few reads map to theY chromosome (sequences were of females), indicating very little contribu-tion from sperm in the reproductive tracts.

Simulation of Reads with Tandem Repeats. For each kmer length (k = 2–10),we generated 600,000 100-bp reads. Each read contained a random numberof tandem occurrences for a randomly generated kmer. One-third of thereads contained perfect tandem repeats, one-third contained 1–4 pointmutations in the tandem repeats, and one-third contained an indel of varyingsize within the tandem repeats. k-Seek was applied to the simulated reads, andcorrect identification for tandem repeats greater than 50 bp was recorded. Wealso generated 200,000 100-bp reads containing random sequences fromwhichno kmers were identified.

Quantifying Satellites with Dot Blots. We radiolabeled 50 pmol of AACA-TAAGATAACATAAGATAACATAAGAT (Sigma) with [γ-32]ATP using T4polynucleotide kinase (NEB), followed by clean-up with Micro Bio-Spin P30Column (Bio-Rad). The probe was denatured at 95 °C for 10 min and im-mediately put on ice. We extracted the DNA of 50 females from 27 lines withthe DNeasy kit (Qiagen). Using a Bio-Dot Microfiltration Apparatus (Bio-Rad), we loaded 100 ng of each sample in 0.4 M NaOH and 10 mM EDTA intriplicate onto a Zeta-Probe GT membrane (Bio-Rad) following the manu-facturer’s instructions, in addition to a threefold serial dilution of DNA fromline B10. The placement of samples was randomized. After drying at 80 °Cfor 30 min, the membrane was incubated in 25 mL of hybridization buffer[0.5 M sodium phosphate, 7% (wt/vol) SDS] with 100 μL of denatured salmonsperm DNA (10 mg/mL) for 30 min at 60 °C in a rotating oven. The buffer wasthen replaced with 25mL of fresh hybridization buffer containing the denaturedprobe and incubated overnight. The membrane was washed at 68 °C twice with50 mL 1× SSC and 0.1% SDS followed by two washes with 0.1× SSC and 0.1%SDS. The membrane was wrapped with plastic wrap, placed into a

Wei et al. PNAS | December 30, 2014 | vol. 111 | no. 52 | 18797

POPU

LATION

BIOLO

GY

Page 6: Correlated variation and population ... - Virginia Tech

phosphoimager for 48 h, and scanned with a Typhoon 9400. Signal intensitywas processed in ImageJ with background subtraction. The intensity of eachsample was calculated according to the standard curve constructed from thedilution series.

In Situ Hybridization. Brains fromwandering third instar larvaewere dissected andwashed in 0.7% NaCl, transferred to 0.5% sodium citrate for 10 min, followed byfixation in 20 μL of 50% acetic acid and 4% paraformaldehyde for 2 min ona siliconized coverslip. The samples were then squashed onto a glass slide and flashfrozen in liquid nitrogen. Slides were then immersed in 100%EtOH for 10min andair-dried at room temperature for 2–3 d with the coverslip removed. Hybridizationprocedure was conducted as in ref. 56. We used 250 ng of AAGAGAAGAGAA-GAGAAGAGAAGAG-Cy3 and AAAATAGAATAAAATAGAATAAAATAGAAT-Cy5probes (Sigma) for the probe mixture. The samples were imaged on a Zeiss con-focal microscope, and images were processed on Zen software.

kmer Correlation Matrix. We applied the publicly available software Modu-lated Modularity Clustering (41) on the normalized counts of the top 100kmers to generate the clustered correlation matrix.

Interspersed kmer Analysis. Using custom Perl scripts on .sep outputs fromk_counter.pl, we identified the number of mate pairs where both readscontain kmers and tallied across all lines to obtain nij , the number of matepairs containing kmer i and kmer j. Interspersed frequency is calculated asnij=

ffiffiffiffiffiffiffiffiffi

ninjp

, where and ni and nj are number of pairs where at least one of thereads contains kmer i and kmer j, respectively.

ACKNOWLEDGMENTS. We thank J. Lis and J. Werner for help with probelabeling, and the Cornell Biotechnology Resource Center Imaging Facility forassistance with the FISH imaging. This work was supported by NationalInstitutes of Health Grants R01 GM074737 (to D.A.B.) and R01 GM64590(to A.G.C.).

1. Doolittle WF, Sapienza C (1980) Selfish genes, the phenotype paradigm and genomeevolution. Nature 284(5757):601–603.

2. Orgel LE, Crick FH (1980) Selfish DNA: The ultimate parasite. Nature 284(5757):604–607.3. Peng JC, Karpen GH (2007) H3K9 methylation and RNA interference regulate nucle-

olar organization and repeated DNA stability. Nat Cell Biol 9(1):25–35.4. Bzymek M, Lovett ST (2001) Instability of repetitive DNA sequences: The role of

replication in multiple mechanisms. Proc Natl Acad Sci USA 98(15):8319–8325.5. Shelby RD, Vafa O, Sullivan KF (1997) Assembly of CENP-A into centromeric chromatin

requires a cooperative array of nucleosomal DNA contact sites. J Cell Biol 136(3):501–513.

6. Malik HS, Henikoff S (2001) Adaptive evolution of Cid, a centromere-specific histonein Drosophila. Genetics 157(3):1293–1298.

7. Karpen GH, Le MH, Le H (1996) Centric heterochromatin and the efficiency ofachiasmate disjunction in Drosophila female meiosis. Science 273(5271):118–122.

8. Henikoff S, Ahmad K, Malik HS (2001) The centromere paradox: Stable inheritancewith rapidly evolving DNA. Science 293(5532):1098–1102.

9. Lohe AR, Roberts PA (2000) Evolution of DNA in heterochromatin: The Drosophilamelanogaster sibling species subgroup as a resource. Genetica 109(1-2):125–130.

10. Kamm A, Galasso I, Schmidt T, Heslop-Harrison JS (1995) Analysis of a repetitive DNAfamily from Arabidopsis arenosa and relationships between Arabidopsis species. PlantMol Biol 27(5):853–862.

11. Lohe AR, Brutlag DL (1986) Multiplicity of satellite DNA sequences in Drosophilamelanogaster. Proc Natl Acad Sci USA 83(3):696–700.

12. Wu C-I, Lyttle TW, Wu M-L, Lin G-F (1988) Association between a satellite DNA se-quence and the Responder of Segregation Distorter in D. melanogaster. Cell 54(2):179–189.

13. Abad JP, et al. (1992) Dodeca satellite: A conserved G+C-rich satellite from the cen-tromeric heterochromatin of Drosophila melanogaster. Proc Natl Acad Sci USA 89(10):4663–4667.

14. Peacock WJ, et al. (1974) The organization of highly repeated DNA sequences in Dro-sophila melanogaster chromosomes. Cold Spring Harb Symp Quant Biol 38:405–416.

15. Lohe AR, Hilliker AJ, Roberts PA (1993) Mapping simple repeated DNA sequences inheterochromatin of Drosophila melanogaster. Genetics 134(4):1149–1174.

16. Lohe AR, Brutlag DL (1987) Identical satellite DNA sequences in sibling species ofDrosophila. J Mol Biol 194(2):161–170.

17. Gall JG, Cohen EH, Polan ML (1971) Reptitive DNA sequences in drosophila. Chro-mosoma 33(3):319–344.

18. Ferree PM, Barbash DA (2009) Species-specific heterochromatin prevents mitoticchromosome segregation to cause hybrid lethality in Drosophila. PLoS Biol 7(10):e1000234.

19. Bayes JJ, Malik HS (2009) Altered heterochromatin binding by a hybrid sterility pro-tein in Drosophila sibling species. Science 326(5959):1538–1541.

20. Satyaki PRV, et al. (2014) The Hmr and Lhr hybrid incompatibility genes suppressa broad range of heterochromatic repeats. PLoS Genet 10(3):e1004240.

21. Smith GP (1976) Evolution of repeated DNA sequences by unequal crossover. Science191(4227):528–535.

22. Okumura K, Kiyama R, Oishi M (1987) Sequence analyses of extrachromosomal Sau3Aand related family DNA: Analysis of recombination in the excision event. Nucleic AcidsRes 15(18):7477–7489.

23. Levinson G, Gutman GA (1987) Slipped-strand mispairing: A major mechanism forDNA sequence evolution. Mol Biol Evol 4(3):203–221.

24. Charlesworth B, Sniegowski P, Stephan W (1994) The evolutionary dynamics of re-petitive DNA in eukaryotes. Nature 371(6494):215–220.

25. Stephan W (1986) Recombination and the evolution of satellite DNA. Genet Res 47(3):167–174.

26. Stephan W, Cho S (1994) Possible role of natural selection in the formation of tan-dem-repetitive noncoding DNA. Genetics 136(1):333–341.

27. Hartl DL (2000) Molecular melodies in high and low C. Nat Rev Genet 1(2):145–149.28. Petrov DA (2001) Evolution of genome size: New approaches to an old problem.

Trends Genet 17(1):23–28.29. Lemos B, Araripe LO, Hartl DL (2008) Polymorphic Y chromosomes harbor cryptic

variation with manifold functional consequences. Science 319(5859):91–93.30. Lemos B, Branco AT, Hartl DL (2010) Epigenetic effects of polymorphic Y chromo-

somes modulate chromatin components, immune response, and sexual conflict. ProcNatl Acad Sci USA 107(36):15826–15831.

31. Malik HS (2009) The centromere-drive hypothesis: A simple basis for centromere

complexity. Prog Mol Subcell Biol 48:33–52.32. Hoskins RA, et al. (2007) Sequence finishing and mapping of Drosophila melanogaster

heterochromatin. Science 316(5831):1625–1628.33. Bosco G, Campbell P, Leiva-Neto JT, Markow TA (2007) Analysis of Drosophila species

genome size and satellite DNA content reveals significant differences among strains

as well as between species. Genetics 177(3):1277–1290.34. Aird D, et al. (2011) Analyzing and minimizing PCR amplification bias in Illumina

sequencing libraries. Genome Biol 12(2):R18.35. Stephan W, Li H (2007) The recent demographic and adaptive history of Drosophila

melanogaster. Heredity (Edinb) 98(2):65–68.36. David JR, Capy P (1988) Genetic variation of Drosophila melanogaster natural pop-

ulations. Trends Genet 4(4):106–111.37. Begun DJ, Aquadro CF (1993) African and North American populations of Drosophila

melanogaster are very different at the DNA level. Nature 365(6446):548–550.38. Schlötterer C, Neumeier H, Sousa C, Nolte V (2006) Highly structured Asian Drosophila

melanogaster populations: A new tool for hitchhiking mapping? Genetics 172(1):

287–292.39. Slatkin M (1995) A measure of population subdivision based on microsatellite allele

frequencies. Genetics 139(1):457–462.40. Hardy OJ, Charbonnel N, Fréville H, Heuertz M (2003) Microsatellite allele sizes: A

simple test to assess their significance on genetic differentiation. Genetics 163(4):

1467–1482.41. Stone EA, Ayroles JF (2009) Modulated modularity clustering as an exploratory tool

for functional genomic inference. PLoS Genet 5(5):e1000479.42. Pool JE, et al. (2012) Population genomics of sub-saharan Drosophila melanogaster:

African diversity and non-African admixture. PLoS Genet 8(12):e1003080.43. Dover G (1982) Molecular drive: A cohesive mode of species evolution. Nature

299(5879):111–117.44. Pollard DA, Iyer VN, Moses AM, Eisen MB (2006) Widespread discordance of gene

trees with species tree in Drosophila: Evidence for incomplete lineage sorting. PLoS

Genet 2(10):e173.45. Csink AK, Henikoff S (1998) Something from nothing: The evolution and utility of

satellite repeats. Trends Genet 14(5):200–204.46. Biggin MD, Tjian R (1988) Transcription factors that activate the Ultrabithorax pro-

moter in developmentally staged extracts. Cell 53(5):699–711.47. Kerrigan LA, Croston GE, Lira LM, Kadonaga JT (1991) Sequence-specific transcrip-

tional antirepression of the Drosophila Krüppel gene by the GAGA factor. J Biol Chem

266(1):574–582.48. Lu Q, Wallrath LL, Granok H, Elgin SC (1993) (CT)n (GA)n repeats and heat shock el-

ements have distinct roles in chromatin structure and transcriptional activation of the

Drosophila hsp26 gene. Mol Cell Biol 13(5):2802–2814.49. Raff JW, Kellum R, Alberts B (1994) The Drosophila GAGA transcription factor is as-

sociated with specific regions of heterochromatin throughout the cell cycle. EMBO J

13(24):5977–5983.50. Platero JS, Csink AK, Quintanilla A, Henikoff S (1998) Changes in chromosomal lo-

calization of heterochromatin-binding proteins during the cell cycle in Drosophila.

J Cell Biol 140(6):1297–1306.51. Pak DT, et al. (1997) Association of the origin recognition complex with hetero-

chromatin and HP1 in higher eukaryotes. Cell 91(3):311–323.52. Rodriguez Alfageme C, Rudkin GT, Cohen LH (1980) Isolation, properties and cellular

distribution of D1, a chromosomal protein of Drosophila. Chromosoma 78(1):1–31.53. Török T, Harvie PD, Buratovich M, Bryant PJ (1997) The product of proliferation dis-

rupter is concentrated at centromeres and required for mitotic chromosome con-

densation and cell proliferation in Drosophila. Genes Dev 11(2):213–225.54. Ellis LL, et al. (2014) Intrapopulation genome size variation in D. melanogaster reflects

life history variation and plasticity. PLoS Genet 10(7):e1004522.55. Greenberg AJ, Hackett SR, Harshman LG, Clark AG (2011) Environmental and

genetic perturbations reveal different networks of metabolic regulation. Mol Syst

Biol 7:563.56. Dernburg AF (2011) In situ hybridization to somatic chromosomes in Drosophila. Cold

Spring Harb Protoc 2011(9):pii: pdb.top065540.

18798 | www.pnas.org/cgi/doi/10.1073/pnas.1421951112 Wei et al.

Page 7: Correlated variation and population ... - Virginia Tech

1

Wei et al; Supporting Information Appendix

Table of contents

Description and usage of k-Seek pg. 2-5

Supplementary table pg. 6

Supplementary figures pg. 7-15

Page 8: Correlated variation and population ... - Virginia Tech

2

Description and usage of the k-Seek package:

The two main programs in the package are k_finder.pl and k_counter.pl, which identifies and

quantifies kmers, respectively.

Command:

perl k_finder.pl input.fastq output

k_finder.pl breaks the ~100 bp sequence reads from a fastq file into small fragments which are

tallied in a hash table indexed by the sequence of the fragments.

For example:

ATCGAATCGAATCGAATCGAATCGAATCGAATCGAATCGAATCGAATCGA

ATCGAATCGAATCGAATCGAATCGAATCGAATCGAATCGAAGGCTAGATG

5mer hash:

Index Count Keep

ATCGA 18 Yes

AGGCT 1 No

AGATG 1 No

To determine the most appropriate fragment length, hash tables are generated for word lengths of

5-10 bps for each read. For each hash table, fragments with very few counts are discarded as they

are largely derived from non-repetitive sequences. The cutoff to discard fragments was adjusted

empirically for optimal performance when processing 100 bp reads. Hash tables with only one

index will be deemed as the correct fragment length.

For example:

Here is ACTGACTGAT repeated 10 times:

ACTGACTGATACTGACTGATACTGACTGATACTGACTGATACTGACTGAT

ACTGACTGATACTGACTGATACTGACTGATACTGACTGATACTGACTGAT

Page 9: Correlated variation and population ... - Virginia Tech

3

5-10mer hash tables are generated from this sequence:

5mer

6mer

7mer

Index Count

Index Count

Index Count

ACTGA 10 ACTGAC 4

ACTGACT 2

CTGAT 10 TGATAC 3

GATACTG 2

TGACTG 3

ACTGATA 2

ATACTG 3

CTGACTG 2

ACTGAT 3

ATACTGA 1

TGAT 1

CTGATAC 1

TGACTGA 1

TACTGAC 1

TGATACT 1

GACTGAT 1

AT 1

8mer

9mer

10mer

Index Count

Index Count

Index Count

ACTGACTG 3 ACTGACTGA 2

ACTGACTGAT 10

ATACTGAC 3 TACTGACTG 1

TGATACTG 2

ATACTGACT 1

ACTGATAC 2

ACTGATAC 1

TGACTGAT 2

GATACTGAC 1

TGAT 1

TGATACTGA 1

CTGATACTG 1

ACTGATACT 1

GACTGATAC 1

TGACTGATA 1

CTGACTGAT 1

T 1

Only the 10mer hash table contains a single unique index, and therefore produces the best

identification.

In cases where the smallest hash table contains two indices, k_finder.pl checks whether

the two indices are offsets of the same sequence (e.g. ACTG, CTGA, TGAC, and GACT). This

will capture kmers interrupted by indels, as they shift the “reading frame” of sequences.

Page 10: Correlated variation and population ... - Virginia Tech

4

After the initial identification process, k_finder.pl checks the sequence for internal

repetition, by further breaking it into substrings. This allows 2-4mers to be identified. For

example ACTGACTG will be identified from the initial 8mer hash. However, this sequence will

be tested for internal repeats, revealing two ACTGs and therefore recorded as a 4mer instead of a

8mer. For each read, the program outputs the fastq format with an additional line providing the

inferred kmer, and an estimate of its occurrence based on its value in the hash table. The output

file here will contain .rep suffix.

Command:

perl k_counter.pl output.rep

The k_counter.pl program takes the inferred kmer and word-searches the fastq read for the

number of occurrences using Perl’s regular expression capability. To ensure tandem-ness, it only

counts search hits that are either immediately preceded or followed by another hit. If two hits are

gapped by a sequence of the same length, k-counter examines that sequence for single bp

differences with the inferred kmer. For 4-10mers, if the gap sequence contains only one bp

difference from the kmer, it is also counted, thus capturing point mutations and sequence errors.

However, gap sequences with more than one bp difference are not counted.

Example:

Here is the 10mer ACTGACTGAT repeated but with an insertion (red) and a substitution (gray

background).

ACTGACTGATACTGACCACCTGATACTGACTGATACTGACTGATACTGAC

TGATACTGACTGATACTGACTGATACTAACTGATACTGACTGATACTGAC

k-finder will identify the 10mer TGATACTGAC repeated 7 times as the following 10mer hash

table will be generated:

Page 11: Correlated variation and population ... - Virginia Tech

5

Index Count Keep

ACTGACTGAT 1 No

ACTGACCACC 1 No

TGATACTGAC 7 Yes

TGATACTAAC 1 No

k-counter word searches the read with TGATACTGAC, counting all matches (underlined).

ACTGACTGATACTGAC1CACCTGATACTGAC

2TGATACTGAC

3TGATACTGAC

4

TGATACTGAC5TGATACTGAC

6TGATACTAACTGATACTGAC

7TGATACTGAC

8

TGATACTAAC is also counted as it contains only one substitution. However, because the first

match is neither immediately preceded nor followed by another copy it is discarded, yielding a

total of 8 counts of the 10mer TGATACTGAC – one more than the counts from k-finder.

k-counter produces two outputs with .sep and .total suffixes. The former is similar to .rep

containing the updated counts with incorrect kmer calls removed. The latter is a list of all

inferred kmers and their total counts across the entire .sep file.

To obtain estimates of the total kmer quantities from multiple WGSs, use k_compiler.pl.

Command:

perl k_compiler.pl folder/ output

The k_compiler.pl program will take a folder containing multiple .total files. It outputs a tab

delimited table where each column is a kmer, and each line is one sequencing sample. In addition,

reverse complements and different offsets of the same kmer (e.g. AAGAG, AGAGA, GAGAA,

AGAAG, GAAGA) are summed; the offset of the highest alphabetical order is listed (e.g.

AAGAG). Table S2 shows the counts of all kmers for the 84 lines, normalized to a bp per 1x

genome sequence (this is why there are fractional counts).

The package is optimized to process 100 bp reads, but can be applied to longer and shorter reads.

Page 12: Correlated variation and population ... - Virginia Tech

6

Supplementary table

Table S1. Counts for all possible kmers and

identified kmers.

Possible

kmers

Median >

100bp Top 100 All

2mer 4 3 3 3

3mer 10 5 6 9

4mer 33 1 1 18

5mer 102 17 21 59

6mer 350 12 14 113

7mer 1170 7 13 115

8mer 4140 4 4 77

9mer 14560 6 8 72

10mer 52632 17 30 202

total 73001 72 100 668

Page 13: Correlated variation and population ... - Virginia Tech

2mer

3mer

4mer

5mer

6mer

7mer

8mer

9mer

10mer

bp

Fre

q o

f corr

ect id

entification

Perfect tandems

Fre

q o

f co

rre

ct id

en

tifica

tion

2mer

3mer

4mer

5mer

6mer

7mer

8mer

9mer

10mer

bp

Substitutions (1-4)

Fre

q o

f co

rre

ct id

en

tifica

tion

2mer

3mer

4mer

5mer

6mer

7mer

8mer

9mer

10mer

bp

Indel

Supplementary Figure S1.

Breakdown of k-Seek

performance. Frequency of

correct identification is plotted for

all kmer sizes, as well as the size

of repetition greater than 50bp for

perfect tandems (A), tandems

with up to 4 substitutions (B), and

tandems with indels (C).

A

B

C

7

Supplementary Figures

Page 14: Correlated variation and population ... - Virginia Tech

Supplementary Figure S2. Correlation of the ZW155 replicates.

replicate1

replic

ate

2

replicate2 re

plic

ate

3

replicate3

replic

ate

1

Pearson’s r = 0.9757

Pearson’s r = 0.9879

Pearson’s r = 0.9816

8

Page 15: Correlated variation and population ... - Virginia Tech

No

rma

lize

d S

ign

al In

ten

sity

Normalized kmer counts

Pearson's

r rep1 rep2 rep3 rep4

rep1 1.000 0.364 0.042 0.411

rep2 0.364 1.000 0.169 0.279

rep3 0.042 0.169 1.000 -0.028

rep4 0.411 0.279 -0.028 1.000

Supplementary Figure S3. Dot Blots with AAGAG probes. A. On each of the 4 membranes

(rep1-4) independently hybridized, 86 different DNA samples were blotted randomly along with

serial dilutions. Pair-wise correlation values of the normalized signal intensity is tabulated. B. 27

lines were randomly blotted in triplicates onto one membrane. Normalized signal intensity

averaged across the triplicates is plotted against the normalized kmer count. No correlation was

found.

A B

9

Page 16: Correlated variation and population ... - Virginia Tech

1

2

3

4

5

6

7

0

Mb

Supplementary Figure S4. Distribution of all kmers across lines.

10

Page 17: Correlated variation and population ... - Virginia Tech

Supplementary Figure S5.

Hierarchical clustering of the

lines. Lines are clustered

based on euclidean distance.

Euclidean distances between

lines were calculated based on

log10 kmer quantities relative to

the mean. The top 100 kmers

were used. The different

populations are color coded:

Beijing (red), Ithaca (green),

Netherlands (blue), Tasmania

(yellow), and Zimbabwe (gray).

N22 B10

ZW142 I13 I31 I16

B59 I06 I26

N07 T23 N18 N01 N11 N15 I35

T22A N16 N23 N14 T07

T43A N04

ZW177 ZW155 ZW185

ZH26 T10

T14A T04 T09 N03 N25 N02 N30 N19

T45B T05 T35 N10

T25A N13 T39

ZH23 T29A

T01 T30A

T24 T36B

I29 ZS10

I23 B51 I33

ZH42 I02 I07

B05 B04 B38 B23 B52 B12 B43 B14 B54 I04

B28 B42

ZH33 ZW144

N17 I01 I24 I03

N29 ZW09

ZW139 ZW140

I17 I38 I34

B11 I22

11

Page 18: Correlated variation and population ... - Virginia Tech

Supplementary Figure S6. Fluorescent in situ hybridization with probes targeting 5mer

AAGAG and population-specific 10mer AACATATAAT.

N22

T23

T23

B04

B14

DAPI AACATATAAT AAGAG Merge

12

Page 19: Correlated variation and population ... - Virginia Tech

-1 0 1 Spearman’s ρ

AAGGAGGAG AGG

AAAATAGAC AATAC

AAAAGAC AAAAC ACATG

AAGACATGAC AATAGAC AAAGAC

AAAACAT AAACAAT

AAGACTAGAC AGATG

AAATTACT AAATGTT

AAGAGTAGAG AAT

ACATATAT AATAATCAT

AATACT AACTTAAT

AGATAT AT

AAAACATAT AACATAT

AATATCAT AATAG

AAGAGCAGAG ACAGAT

AACATATAAT AATAT

AAAATAGAAT AAAATAT

ACCAGATAG ACTG

AGATGG AAGGAG

ACAGACCAG AAAATAACAT AACACACAC

AC ACACG

AACACAC ACGAG

AAGCAC ACATAG

AACACAAGAC AAGAGATGAG

AACAC AAACAC AGAGC

AAC AAGAC ACTCT

AAGACAGGAC AAGTG

AAAAGAG AAGAGAAGTG

AATGT AAGGAT AAGGC

AAGGCAATGC AACAG ACAGG AGAGG

AAGACAAGAG AAAAGAGAAG AAGACCAGAC

AAAGAG AAGAGG

AAGACAAGAT AAGACAAGGC

AAAAG AAGAT

AG AAGAG

AAGAGAG AAGAGAGAG

AAGAGAAGGG AAGAGACGAG

ACTGGG AGC

AAGAGGAGAG AAG

AACAGAG AAGAGAAGCG AAGAGAATAG AACATAGAAT AACATTGAAT AAAGAATAAC

AAAAGG AAGAGAAGAT

AGAT ACT

AAATGAATG AAGAGACGGG

AATTATC AAAAGGAG

ACATACATAT

AAGGAGGAG

AGG

AAAATAGAC

AATAC

AAAAGAC

AAAAC

ACATG

AAGACATGAC

AATAGAC

AAAGAC

AAAACAT

AAACAAT

AAGACTAGAC

AGATG

AAATTACT

AAATGTT

AAGAGTAGAG

AAT

ACATATAT

AATAATCAT

AATACT

AACTTAAT

AGATAT

AT

AAAACATAT

AACATAT

AATATCAT

AATAG

AAGAGCAGAG

ACAGAT

AACATATAAT

AATAT

AAAATAGAAT

AAAATAT

ACCAGATAG

ACTG

AGATGG

AAGGAG

ACAGACCAG

AAAATAACAT

AACACACAC

AC

ACACG

AACACAC

ACGAG

AAGCAC

ACATAG

AACACAAGAC

AAGAGATGAG

AACAC

AAACAC

AGAGC

AAC

AAGAC

ACTCT

AAGACAGGAC

AAGTG

AAAAGAG

AAGAGAAGTG

AATGT

AAGGAT

AAGGC

AAGGCAATGC

AACAG

ACAGG

AGAGG

AAGACAAGAG

AAAAGAGAAG

AAGACCAGAC

AAAGAG

AAGAGG

AAGACAAGAT

AAGACAAGGC

AAAAG

AAGAT

AG

AAGAG

AAGAGAG

AAGAGAGAG

AAGAGAAGGG

AAGAGACGAG

ACTGGG

AGC

AAGAGGAGAG

AAG

AACAGAG

AAGAGAAGCG

AAGAGAATAG

AACATAGAAT

AACATTGAAT

AAAGAATAAC

AAAAGG

AAGAGAAGAT

AGAT

ACT

AAATGAATG

AAGAGACGGG

AATTATC

AAAAGGAG

ACATACATAT

Supplementary Figure S7. Clustered pairwise-correlation matrix of kmers across

lines. Same as Fig. 4A but with all kmers labeled.

13

Page 20: Correlated variation and population ... - Virginia Tech

Spearman’s ρ = -0.270, p < 0.01

Pearson’s r = -0.295, p < 0.05

AACATATAAT

ACCAGATAG AAGAGCAGAG

AAT

Spearman’s ρ = -0.358, p < 0.001

Pearson's r = -0.256, p < 0.05

Supplementary Figure S8. Anti-correlated kmers. Spearman’s ρ is calculated with all

samples. Pearson’s r is calculated after removing samples that have kmer quantities lower

than 1% of the average. Red line depicts regression line using all samples.

14

Page 21: Correlated variation and population ... - Virginia Tech

Co

eff

icie

nt o

f va

riatio

n

Supplementary Figure S9. Distribution of coefficient of variation.

The coefficient of variation for the top 100 kmers are plotted. The

Zimbabwe population has the highest average coefficient of variation

(p = 0.0244, Kruskal Wallis test.)

15