MECHANISMS OF ADAPTATION AND SPECIATION: AN EXPERIMENTAL STUDY USING ARTIFICIAL LIFE By Carlos Jesus Anderson A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Zoology – Doctor of Philosophy Ecology, Evolutionary Biology and Behavior – Dual Major 2013 ABSTRACT MECHANISMS OF ADAPTATION AND SPECIATION: AN EXPERIMENTAL STUDY USING ARTIFICIAL LIFE By Carlos Jesus Anderson Detailed experimental studies in evolutionary biology are sometimes difficult—even with model organisms. Theoretical models alleviate some of these difficulties and often provide clean results, but they cannot always capture the complexity of dynamic evolutionary processes. Artificial life systems are tools that fall somewhere between model organisms and theoretical models that have been successfully used to study evolutionary biology. These systems simulate simple organisms that replicate, acquire random mutations, and reproduce differentially; as a consequence, they evolve naturally (i.e., evolution itself is not simulated). Here I use the software Avida to study several open questions on the genetic mechanisms of adaptation and speciation. In Chapter 1 (p. 15), I investigated whether beneficial alleles during adaptation came from new mutations or standing genetic variation—alleles already present in the population. I found that most beneficial alleles came from standing genetic variation, but new mutations were necessary for long-term evolution. I also found that adaptation from standing genetic variation was faster than from new mutations. Finally, I found that recombination brought together beneficial combinations of alleles from standing genetic variation. In Chapter 2 (p. 35), I investigated the probability of compensatory adaptation vs. reversion. Compensatory adaptation is the fixation of mutations that ameliorate the effects of deleterious mutations while the original deleterious mutations remain fixed. I found that compensatory adaptation was very common, but the window of opportunity for reversion was increased when the initial fitness of the population was high, the population size was large, and the mutation rate was high. The reason that the window of opportunity for reversion was constrained was that negative epistatic interactions with compensatory mutations prevented the revertant from being beneficial to the population. In Chapter 3 (p. 66), I showed experimentally that compensatory adaptation can lead to reproductive isolation (specifically, postzygotic isolation). In addition, I found that the strength of this isolation was independent of the effect size of the original deleterious mutations. Finally, I found that both deleterious and compensatory mutations contribute equally to reproductive isolation. Reproductive isolation between populations often evolves as a byproduct of independent adaptation to new environments, but the selective pressures of these environments may be divergent (‘ecological speciation’) or uniform (‘mutation-order speciation’). In Chapter 4 (p. 86), I compared directly the strength of postzygotic isolation generated by ecological and mutation-order processes with and without migration. I found that ecological speciation generally formed stronger isolation than mutation-order speciation and that mutation-order speciation was more sensitive to migration than ecological speciation. Under the Dobzhansky-Muller model of speciation, hybrid inviability or sterility results from the evolution of genetic incompatibilities (DMIs) between species-specific alleles. This model predicts that the number of pairwise DMIs between species should increase quadratically through time, but the few tests of this ‘snowball effect’ have had conflicting results. In Chapter 5 (p. 115), I show that pairwise DMIs accumulated quadratically, supporting the snowball effect. I found that more complex genetic interactions involved alleles that rescued pairwise incompatibilities, explaining the discrepancy between the expected accumulations of DMIs and observation. To the memory of Harry Lee Moore “Mr. Moore” (January 23, 1955–March 17, 2012) iv ACKNOWLEDGMENTS I would like to thank my advisor, Dr. Barry Williams, for taking me under his mentorship. Barry has been an excellent source of ideas that have helped me frame and focus my dissertation. He has been very supportive throughout my time here and open to my exploration of new topics, some of which turned out to be chapters in my dissertation. Through the professional connections he has helped me make, Barry has helped me further my scientific career. Finally, Barry has been a friend as well as a colleague, and through him I have learned much about life. I would like to thank my dissertation committee, Drs. Richard Lenski, Charles Ofria, and Doug Schemske. They have been incredibly helpful in developing my ideas and focusing my research. Listening to them discuss science in my committee meetings was always inspiring. I would also like to thank Dr. Luke Harmon, who collaborated with me on one of my chapters. Luke has been very positive and cared for my well-being because he usually started our conversations with “Are you having fun?” I would like to thank my other collaborators, who, by sharing their knowledge and wisdom, have made me a better scientist: Dr. Jenny Boughman, Dr. Alex Shingleton, Dr. Jeff Barrick, Dr. Shampa Ghosh, Dr. Lili Lettieri, Dr. Jason Keagy, Dr. Bjørn Østman, and Alycia Lackey. I would like to thank the many excellent people of the Williams lab—past and present. Vicki Rademacher, Kristine Cox, Krista Reitenga, and Siva Nallasivam did a lot of the lab work that some of my research explorations relied upon. Alex Prediger, Bahaar Chawla, Molly Roseland, and other undergraduates helped me run experiments or kept the lab clean and well-supplied. I thank my fellow graduate lab members, Krista Reitenga, Katy Califf, Chris Hamm, Cory Kohn, and Michael Grispo. v I would like to thank the Devolab for their great feedback during lab presentations that have greatly improved my research. Several people (including friends named later) offered great comments on parts of my work, especially Dr. Ian Dworkin, Dr. Aleeza Gerstein, Dr. Daniel Matute, Emily Weigel, and Dr. Sonal Singhal. I would also like to thank Dr. Danita Brandt, Dr. Michael Velbel, Dr. Michael Gottfried, Dr. Robert Anstey, and Summer Ostrowski for enriching my mind with great scientific discussions. I would like to thank the many people who worked in the background and made my life a little easier: Brian Baer, Pat Bills, Dirk Colbry, Rob Lyon, Lisa Craft, Julia Ahmed, A Pat Resler, and Connie James. Dr. Clifford Weil taught me L TEX and provided me with the template that helped create this beautiful-looking dissertation. My work would have been impossible without the many agencies and organizations that funded me: Michigan State University, the Department of Zoology, the Quantitative Biology Initiative, the Graduate School, the Ecology, Evolutionary Biology and Behavior (EEBB) Program, and the BEACON Center for the Study of Evolution in Action. Throughout my six years here, I have made some amazing friends, which alone has made this experience worthwhile. I especially thank Eli Swanson, Austin Dreyer, and Nick Testa. I would also like to offer warm thanks to Greg Stricker, Kelly Barnes, Chelsea Van Eck, Darby Monks, Aaron Florn, Ben Dantzer, Adrian Dantzer, Emily Dittmar, Katy Califf, Heather Stadden, Gaurav Moghe, Lorna Watt, Mridul Thomas, Kane Keller, and Lori TorgersonWhite. Back home, I would like to thank my best friends, Enrique Pereira and Duvan Cope, and my high school physics teacher “Mr. Moore,” who was an inspiration to me and to whom I dedicate this dissertation. Finally, I would like to thank my family. My parents, Jesus and Crystel, my sister Cristel, and my brothers Pablo and Marcelo have given me love, support, and encouragevi ment throughout my life. I would also like to thank my grandparents and extended family, especially my grandmother Zarela, who will soon turn 93, for sharing her love and wisdom with me. Finally, I would like to thank my little companions, both past and present, who have done nothing but to simply be cute, especially Baxter, Johnny, and Angel. vii TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii Introduction . . . . . . Background . . . . . . Study system: Avida . Dissertation summary LITERATURE CITED . . . . . 1 1 6 9 12 Chapter 1 The role of standing genetic variation in adaptation to a new environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Standing genetic variation in digital organisms . . . . . . . . . . . . . . . . . 1.3 Source of beneficial alleles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Speed of adaptation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5 Fitness effect of random alleles from different sources of variation . . . . . . 1.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . LITERATURE CITED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 15 16 18 23 24 27 32 Chapter 2 The probability of compensatory adaptation 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Frequency of compensatory mutations . . . . . 2.2.2 Effect of initial fitness of mutant . . . . . . . . . 2.2.3 Effect of population size . . . . . . . . . . . . . 2.2.4 Effect of mutation rate . . . . . . . . . . . . . . 2.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.1 Availability of compensatory mutations . . . . . 2.3.2 Effect of initial mutant fitness . . . . . . . . . . 2.3.3 Population size . . . . . . . . . . . . . . . . . . 2.3.4 Mutation rate . . . . . . . . . . . . . . . . . . . 2.3.5 Conclusions . . . . . . . . . . . . . . . . . . . . 2.4 Materials and Methods . . . . . . . . . . . . . . . . . . 2.4.1 Evolution of ancestors . . . . . . . . . . . . . . 2.4.2 Construction and evolution of mutants . . . . . 2.4.3 Detection of compensatory adaptation . . . . . 35 35 38 39 41 48 50 52 54 55 56 57 57 58 58 59 59 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . in digital organisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.4 Estimation of the expected time for reversion . . . . . . . . . . . . . 2.5 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . LITERATURE CITED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 62 63 Chapter 3 Compensatory adaptation causes rapid incipient speciation . . 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 Reproductive isolation via compensation is rapid . . . . . . . . . . . 3.2.2 Compensatory adaptation forms strong genetic incompatibilities . . . 3.2.3 Both deleterious and compensatory mutations contribute to reproductive isolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1 Strains and experimental conditions . . . . . . . . . . . . . . . . . . . 3.4.2 Isolation of mutants . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.3 Compensated populations . . . . . . . . . . . . . . . . . . . . . . . . 3.4.4 Hybridization method . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.5 Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . LITERATURE CITED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 66 69 70 72 Chapter 4 Ecological and mutation-order speciation 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . 4.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1 Experimental design . . . . . . . . . . . . . 4.2.2 Postzygotic reproductive isolation . . . . . . 4.2.3 Genetic divergence . . . . . . . . . . . . . . 4.2.4 Hybrid phenotypes . . . . . . . . . . . . . . 4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.1 Postzygotic reproductive isolation . . . . . . 4.3.2 Genetic divergence . . . . . . . . . . . . . . 4.3.3 Hybrid phenotypes . . . . . . . . . . . . . . 4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . LITERATURE CITED . . . . . . . . . . . . . . . . . . . in digital organisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 5 Experimental evolution of the snowball effect shows the tance of complex epistasis in speciation . . . . . . . . . . 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2.1 Pairwise DMIs increased quadratically through time . . . . . 5.2.2 Hybrid inviability increased linearly through time . . . . . . 5.2.3 Secondary alleles rescue pairwise DMIs . . . . . . . . . . . . 5.2.4 Derived-ancestral DMIs occur as often as derived-derived . . 5.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix 73 75 78 78 79 80 80 81 81 82 . 86 . 86 . 89 . 89 . 91 . 92 . 93 . 96 . 96 . 99 . 100 . 102 . 111 impor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 115 118 119 120 123 124 125 127 5.4.1 Experimental configuration 5.4.2 Statistics . . . . . . . . . . . 5.5 Acknowledgments . . . . . . . . . . LITERATURE CITED . . . . . . . . . . . . . . x . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 128 128 129 LIST OF TABLES Table 1.1 Table 3.1 Table 4.1 The number of single mutants (out of 1,000), categorized by the mutation’s source and fitness effect (w): lethal (w = 0), strongly deleterious (0 < w ≤ 0.99), mildly deleterious (0.99 < w ≤ 0.999), neutral or nearly neutral (0.999 < w ≤ 1.001), mildly beneficial (1.001 < w ≤ 1.01), and strongly beneficial (w > 1.01). . . . . . . . 26 Number of mutants and compensated populations per ancestor for the different treatments . . . . . . . . . . . . . . . . . . . . . . . . . 70 Genetic divergence (FST ) between demes for each treatment. . . . . 100 xi LIST OF FIGURES Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 2.1 The heterozygosity of each locus of the population before the environmental change. Heterozygosities above 0.095 indicate the presence of standing genetic variation. . . . . . . . . . . . . . . . . . . . . . . . 18 The frequencies of alleles through time for two loci in which an allele became beneficial and subsequently fixed. In the top plot, the beneficial allele came from standing genetic variation, and in the bottom plot, the beneficial allele appeared as a new mutation. Different alleles are represented by different colors. The y-axis in each plot ranges from 0.0 to 1.0. For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation. . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 The cumulative frequency of fixed alleles that became beneficial at a specific frequency (0.05 bin size) for populations that started with standing genetic variation (solid lines) and for control, isogenic populations (dashed lines). The gray lines indicate the 95% bootstrap confidence interval around the mean of 20 replicate populations. The red vertical line indicates the frequency below which alleles were considered to appear as new mutations. The red horizontal lines indicate the proportions of alleles that came from new mutations for either type of population. . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 The mean fitnesses (relative to the ancestor) of populations evolved after an environmental change at (A) 0.1, (B) 0.01, and (C) 0.0 mutations per genome per generation (U ). Populations evolved starting either with the ancestral population (solid line), which contained standing genetic variation (SGV) or with an isogenic population based on the consensus sequence of the ancestral population (dashed line). Gray lines represent the 95% bootstrap confidence intervals around the mean. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 Proportion of one-step mutations that are compensatory. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are the 95% confidence interval of the mean. . . . . . . . . . . . . . . . . . . 40 xii Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 Figure 2.7 Effect of initial fitness on compensation. Top figures: The proportion of runs that fully or partially compensated for (A) ancestor 1 (line is best fit linear model, P = 0.002) and (B) ancestor 2 (P = 0.043). Middle figures: The proportion of runs that reverted for (C) ancestor 1 (P < 0.001) and (D) ancestor 2 (P = 0.072). Bottom figures: The proportional increase in fitness for compensatory runs for (E) ancestor 1 and (F) ancestor 2. . . . . . . . . . . . . . . . . . 42 Generation at which reversion stopped being beneficial at various initial mutant fitness effects. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are the 95% confidence interval of the mean. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 Substitution at which reversion stopped being beneficial. at various initial mutant fitness effects. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Lines are linear regressions for each mutant replicate. Error bars are the 95% confidence interval of the mean. . 46 Generation at which reversion must appear in order to fix. (A) Ancestor 1, Mutant 1, (B) Ancestor 1, Mutant 2, (C) Ancestor 2, Mutant 1, and (D) Ancestor 2, Mutant 2. Open circles ( ) represent the first mutant replicate while solid triangles ( ) represent the second mutant replicate. Error bars are the 95% confidence interval of the mean. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 Effect of population size on compensation. (A) The proportion of runs that underwent compensatory evolution for ancestor 1 and (B) ancestor 2. (C) The proportion of runs that reverted for ancestor 1 and (D) ancestor 2. (E) The proportional increase in fitness for compensatory runs for ancestor 1 and (F) ancestor 2. Open circles ( ) represent the first mutant replicate while solid triangles ( ) represent the second mutant replicate. . . . . . . . . . . . . . . . . . 49 Effect of mutation rate on compensation, reversion, and final population fitness. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are bootstrap 95% confidence intervals of the mean. . . . . . . . . . . . . . . . . . . . . . . . . . . 51 xiii Figure 2.8 Figure 3.1 Figure 3.2 Figure 3.3 Figure 4.1 Generation at which reversion stopped being beneficial at various mutation rates. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are bootstrap 95% confidence intervals of the mean. 53 Fitness of hybrids after 25,000 updates of parental evolution. Populations compensated for either small-effect or large-effect deleterious mutations. (A) Hybridizations between compensated genotypes and their ancestor. (B) Hybridizations between pairs of compensated genotypes sharing the same ancestor. ‘Del. Alone’ include only the initial deleterious mutations before compensation, ‘Comp. Alone’ include only the compensatory mutations after compensation, and ‘Del. + Comp.’ include both deleterious and compensatory mutations. ‘Control’ include all mutations accumulated neutrally. Error bars are 95% bootstrap confidence intervals. . . . . . . . . . . . . . 71 Strength of genetic incompatibilities in hybrids. (A) Least squares linear fit of the mean number of mutations in hybrids between ancestor and compensated genotypes (AC) against their mean fitness. (B) Same as (A) except for hybrids between pairs of compensated genotypes (CC). Insets: Means and 95% bootstrap confidence intervals of each slope where shared letters indicate overlapping confidence intervals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 Fitness of hybrids with intermediate parental contributions of deleterious and compensatory mutations. (A) Mean fitness of hybrids between the ancestor and compensated parents that inherit the specified percent of small-effect deleterious and compensatory mutations. (B) Same as (A) except for large-effect deleterious mutations. . . . . 76 Method to determine whether a hybrid contains genetic incompatibilities. A hybrid is composed of two parental components, C1 and C2 . Note that a parental component is only the parental region inherited by the hybrid; it is not the complete parent. If the hybrid cannot perform the task but either parental component can, then there must be at least one incompatibility between the components. . . . . . . . 95 xiv Figure 4.2 Mean fitness of hybrids (y-axis) between populations evolved under various migration rates (x-axis) and different treatments (markers and lines): the ancestral environment (‘Genetic drift’), different new environments (‘Ecological’), or the same new environment (‘Mutationorder 1’ and ‘Mutation-order 2’). At low dimensionality (A and C), the new environments rewarded only two tasks; at high dimensionality (B and D), the new environments rewarded 34 tasks. With two-point crossover (A and B), only one region was exchanged when creating hybrids; with multiple-point crossover (C and D), up to 100 regions were exchanged. . . . . . . . . . . . . . . . . . . . . . . . . . 97 Figure 4.3 Consensus sequences of the first five evolved replicate pairs of demes in the mutation-order 1 treatment under zero migration (top) and 0.01 migration (bottom). Similar results were observed for the mutation-order 2 treatment. Sequences were 200 instructions in length, but only the loci that differed among each set of five replicates are shown. Derived alleles involved in performing a task are highlighted (black highlight = task 1, gray highlight = task 2, bold font = both). 101 Figure 4.4 Number of hybrids able to perform certain tasks (see Methods) for the (A) ecological, (B) mutation-order-1, and (C) mutation-order-2 treatments under zero migration, low dimensionality, and two-point crossover. Each point is a hybrid count (out of 1,000) for a single replicate. Counts from the same replicate are connected by gray lines. The mean hybrid count per category among replicates is indicated by a horizontal bar. Non-significant differences between the mean hybrid counts share the same letter above (or below) the points in each category. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 Figure 4.5 Number of fit hybrids (i.e., can perform two tasks) whose parental components can perform certain tasks (see Methods) for the (A) mutation-order 1 and (B) mutation-order 2 treatments under zero migration, low dimensionality, and two-point crossover. Each point is a hybrid count for a single replicate. Counts from the same replicate are connected by gray lines. The mean hybrid count per category among replicates is indicated by a horizontal bar. Non-significant differences between the mean hybrid counts share the same letter above the points in each category. . . . . . . . . . . . . . . . . . . . . . . . 104 xv Figure 5.1 The Dobzhansky-Muller model of postzygotic isolation. A population with haploid genotype abcde (bottom) becomes geographically divided into two, and each population independently evolves through time (thick arrows, time progresses upward). In the first population, allele a is substituted with allele A, while in the second population allele b is substituted with allele B. At this point, if the populations were to come into contact and hybridize, their hybrids would have genotypes Abcde, aBcde, or ABcde. Because alleles A and B evolved independently and thus may only function properly in the background in which they evolved, hybrids with genotype ABcde may be sterile or inviable. In this case, alleles A and B are said to form a Dobzhansky-Muller incompatibility (DMI) (arrow from B to A). The populations could have instead hybridized after two or more alleles have fixed within each lineage (e.g., alleles A, C, and E in the first population and alleles B and D in the second), which would have resulted in hybrids with many more possible DMIs. DMIs between derived alleles (‘derived-derived’ DMIs) are thin solid arrows, and DMIs between derived and ancestral alleles (‘derived-ancestral’ DMIs) are thin dashed arrows. (Modified from (Orr, 1995).) . . . . . 116 Figure 5.2 Illustration of the method for identifying (A) derived-ancestral and (B) derived-derived pairwise DMIs (see text). (A) Step 1: I tested the individual fitness effect of each derived allele (black bars) of a species on the ancestral background. Step 2: For each derived allele that was incompatible with the ancestral background (indicated by the black triangle pointing down), I tested the individual effect of the other derived alleles on that genetic background. Step 3: If a second derived allele resulted in high fitness (indicated by the gray triangle pointing up), the ancestral allele at that second locus was incompatible with the original derived allele. (B) Step 1: I tested the individual fitness effect of each derived allele (gray bars) of a species on the ancestral background. Step 2: For each derived allele that was compatible with the ancestral background, I tested the individual effect of each derived allele (itself compatible with the ancestral background) of the other species. Step 3: If the second allele lowered the fitness, then these two allele must be incompatible. . . . . . . . . . . . . . . . . . 121 xvi Figure 5.3 (A) Mean number of pairwise DMIs through time. Each point represents the mean of 20 replicate runs. The black curve represents the quadratic model of the data with parameters estimated using maximum likelihood. The gray lines represent the bootstrap 95% confidence intervals of the means. (B) Mean number of unfit hybrids (out of 10,000) between populations through time. Each point represents the mean of 20 replicate runs. The black line represents the linear model of the data with parameters estimated using maximum likelihood. The gray lines represent the bootstrap 95% confidence intervals of the means. . . . . . . . . . . . . . . . . . . . . . . . . . 122 Figure 5.4 Mean number of all hybrids ( ) and unfit hybrids ( ) (out of 10,000) with at least one pairwise DMI through time. Each point represents the mean of 20 replicate runs. The gray lines represent the bootstrap 95% confidence intervals of the means. . . . . . . . . . . . . . . . . . 124 xvii Introduction Background Richard Feynman once said, “If our small minds, for some convenience, divide this . . . universe, into parts—physics, biology, geology, astronomy, psychology, and so on—remember that nature does not know it.” ‘Biology’ is nothing more than a convenient term used to categorize certain kinds of complex natural patterns and processes. Many of these processes, however, are not phenomena that strictly occur in living systems. For example, the process of diffusion in the electron transport chain, which generates energy for cells in the form of ATP molecules, is used by biological systems but is not exclusively a biological process. Diffusion is a natural phenomenon that occurs whenever particles exhibit Brownian, or random, motion, and was, in fact, first studied by physical scientists. In the same way, evolution via natural selection occurs in biological systems, but it is not exclusively a biological process (Pennock, 2007). Cultural phenomena, such as languages and ideas, and artificial genetic systems, which I discuss later, can also evolve via selection. Natural selection is a universal process that occurs whenever three conditions are met: (1) inheritance, (2) variation, and (3) differential survival/reproduction (Adami, 2006). 1 Therefore, the process of evolution can be studied in isolation. Mathematical models and computer simulations of evolutionary processes show that it can be studied outside of the biological realm. Using such models I have learned a great deal about evolution, creating predictions that can then be tested in biological organisms. However, models are limited in that they cannot capture the complexity and open-endedness of evolution in action (Yedid and Bell, 2001). For example, models often deal with only a range of parameters values and their possible outcomes is often limited in scope, so that new patterns or new behavior is not possible (Yedid and Bell, 2001). Biological models, such as the use of E. coli and S. cerevisiae, have provided great insights about the working of evolution, but even these systems can be intractable. A middle-ground between these extremes—mathematical models and biological ones—is the use of artificial life systems (Yedid and Bell, 2001). These are systems in which the conditions for evolution are present, and therefore constitute an instance of evolution (Pennock, 2007). In my dissertation, I use the artificial life system Avida (Ofria and Wilke, 2004) to study specific genetic mechanisms of adaptation and speciation. Digital populations in Avida meet the conditions required for evolution; therefore, Avida represents an instance of evolution. I discuss how Avida works in more detail in the next section, but briefly, digital organisms consist of a sequence of instructions (a ‘genome’) that is passed on to offspring during replication (meets inheritance condition). Phenotypic and genotypic variation is introduced through mutations and, in sexual populations, through recombination. Organisms with phenotypes that match their environment are able to reproduce faster than others (meets differential survival/reproduction condition). Avida is unlike a mathematical or computational model in that it need not be constrained by parameters and their ranges; evolving digital organisms evolve novelty and complexity that is impossible to have been predicted (Yedid and 2 Bell, 2001; Wilke and Adami, 2002). (Note: Yedid and Bell (2001) investigated the Terra system, from which Avida was based, but I will sometimes cite them when discussing Avida if it is appropriate.) Indeed, the patterns observed in Avida runs can be so complex as to require traditional models in order to understand them (B. Østman, pers. comm.) Avida thus provides an independent system in which predictions made from evolutionary theory could be tested (Yedid and Bell, 2001), and in turn provide feedback to those theories and help refine them (Wilke and Adami, 2002). Apart from providing open-ended evolution and potentially novel patterns, Avida has other advantages due to its computational nature. In Avida, one has complete and accurate knowledge of individual organisms, their genotype, fitness, and lineage, and one can track individual mutations. This power allows one to create experiments of unprecedented sophistication that would have been extremely difficult or even impossible to carry out in biological organisms (Elena and Lenski, 2003). For example, one is able to go back to any point in time after an experiment, change the value of a variable, and re-run the experiment in exactly the same way except for the altered variable. Another advantage is the capacity to run exact replicates of an experimental configuration (not counting the initial random ‘seed,’ of course), providing high statistical power to data analyses. Using Avida, one is able to run experiments for millions of generations in a matter of days or weeks. These benefits are not without a cost. The most prominent is that one is restricted to the kinds of questions to which Avida is best suited (Wilke and Adami, 2002). For example, it would not make sense to study the evolution of mitosis (Avida has no true cell division). Thus, the questions asked should be amenable to abstraction (Wilke and Adami, 2002). Additionally, one may sometimes ‘miss the forest for the trees’ when reporting or interpreting Avida results, particularly when the purpose is to study biological phenomena. Special care 3 must be taken to ensure that observations are not merely artifacts of the computational system in order to make the appropriate biological inferences. Being a computational system, Avida requires a certain level of computer proficiency, especially when performing complex data analyses or when customizations to the system itself are needed. Physiologically, digital organisms in Avida work very differently from biological organisms (Yedid and Bell, 2001), and they are not as complex as even the simplest lifeforms (Lenski, 2001). However, several evolutionary properties have been found to be remarkably similar to that of biological organisms (Yedid and Bell, 2001; Wilke and Adami, 2002; Adami, 2006) (e.g., the distribution of mutational effects, the types of epistasis, and the genetic architecture of sexual organisms). But as R. Lenski points out in O’Neill (2003), “even if the digital and biological realms sometimes come into scientific conflict, it would only lead one to ask why and then probe the relevant factors more deeply.” Therefore, Avida provides an additional avenue of inquiry where experiments can be conducted, and the results compared to those of model organisms and theory, in order to discover the generality of some phenomenon (Wilke and Adami, 2002). In addition to all the benefits that Avida offers to biologists, Avida is also of interest to engineers. Other evolutionary computational tools, such as genetic algorithms, have been used by engineers to tackle difficult problems that are best solved by evolution rather than by design (McKinley et al., 2008). As in applications to biological questions, Avida offers a more open-ended approach to optimization and algorithms for solving engineering problems (McKinley et al., 2008). Evolutionary concepts, such as robustness, evolvability, and cooperation, are very interesting to system and software engineers, who would like to develop systems that can compensate for failures, be resilient to varying parameters, protect themselves from attacks, gather information cooperatively, and be efficient at distributing 4 data through a network (Beckmann et al., 2007; McKinley et al., 2008; Goldsby and Cheng, 2008; Knoester et al., 2009; Knoester and McKinley, 2011). Wilke and Adami (2002) posit that “robots, and the software that directs them, might evolve without human interaction, at which point they would become part of the ecosystem in which I live.” Using Avida, biologists themselves can therefore contribute to solving problems outside their field, broadening their impact of their research (O’Neill, 2003). In this dissertation, I use Avida to investigate some genetic mechanisms of adaptation and speciation. Adaptation is the process in which beneficial traits are acquired by organisms in a population through time. Adaptive traits driven to fixation by natural selection are ultimately encoded in the genetic material of organisms, and therefore adaptation is often thought of in terms of adaptive alleles. One of the goals of research in evolutionary biology is to understand the genetic basis of adaptation. For example, does the raw material for adaptive evolution come from from new mutations or from “standing genetic variation” (i.e., allelic variation present in the population) (Orr, 2005)? Questions like this require detailed analyses of genetic data, such as tracking individual alleles through time (Barrett and Schluter, 2008). Speciation, the process by which new species form, is often a by-product of adaptation (Coyne and Orr, 2004; Schluter, 2009; Sobel et al., 2010). Populations that evolve independently are likely to adapt in different ways and therefore diverge genetically. Genetic differences are likely to create ecogeographic, morphological, physiological, or genetic isolation between populations (Coyne and Orr, 2004; Schemske, 2010), collectively known as “reproductive isolation.” For example, populations may not recognize each other as potential mates or may produce sterile or inviable hybrids. A topic of much recent consideration is the genetic mechanisms that lead to reproductive isolation. Like in studies of the genetic basis 5 of adaptation, an understanding of the genetic basis of speciation requires detailed analyses of genetic data over time. The topic of speciation, as commonly defined as reproductive isolation, has never before been studied in Avida. This was because sexual reproduction in Avida has only recently been implemented, originally to study the evolution of sex (Misevic et al., 2004, 2006, 2010). Sexual reproduction in Avida was implemented by exchanging genetic material between the next two organisms that are ready to reproduce. Currently, there is no mate recognition in Avida (although research in this topic has been done), and therefore prezygotic isolation is not easy to study. My research on speciation has thus focused on postzygotic isolation, i.e., hybrid sterility or inviability, as creating hybrids between organisms and measuring their fitness is simple. My research has demonstrated that Avida is a useful tool to complement other approaches in speciation research as it allows for the direct observation of evolution and reproductive isolation in action. Study system: Avida In this section, I provide a brief overview of Avida; for a full description, see Ofria and Wilke (2004). Avida is freely available at http://avida.devosoft.org. In Avida, digital organisms are composed of a linear sequence of instructions (akin to a haploid genome), memory space in the form of registers and stacks, pointers to memory locations, and a central processing unit (CPU) that executes instructions. The instruction set makes up an assembly-like programming language, consisting of instructions for arithmetic operations, memory manipulation (e.g., swapping registers or pushing into a stack), conditional execution, iteration, input/output operations, and allocation and copying of memory. Organisms execute their 6 instructions sequentially, sometimes skipping instructions for conditional statements or repeating the same instructions inside a loop; when the last instruction is executed, execution starts again at the first instruction. By executing instructions in their genomes, organisms are able to (1) replicate and (2) perform computational ‘tasks’ that increase the speed at which they replicate and thus increase fitness. To replicate, an allocation instruction creates the memory space required by the organism’s offspring, and a copy instruction inside a loop allows the organism to copy itself into the new memory space. The copy instruction that allows organisms to replicate has a configurable probability of making mistakes, which introduces various kinds of mutations. By default, replication is asexual. However, Avida may be configured to perform sexual replication, in which the genomes of two asexually-produced offspring are recombined by exchanging two randomly-sized regions of their genomes. The offspring (whether clonal or recombinants) are put into the population in random locations, replacing whatever organisms were already there. Generations are therefore overlapping, as offspring are born continuously, replacing older individuals but who are likely not their parents. In addition to replication, genomic instructions allow organisms to acquire 32-bit input values and use them to perform computational tasks. Tasks are boolean operations, such as NOT, AND, and OR, and are applied to input values bit by bit. For example, if input values were 8 bits, the operation 10011101 AND 11101011 would produce 10001001 according to the rules of boolean logic for AND (0 AND 0 = 0, 0 AND 1 = 0, 1 AND 0 = 0, and 1 AND 1 = 1. In Avida, however, there is no AND operation nor any other boolean operation except for NAND, from which all other boolean operations may be built, a property of NAND known as ‘functional completeness’ in boolean algebra. For example, P AND Q, where P and Q are input values, is equivalent to (P NAND Q) NAND (P NAND Q). Therefore, in order 7 to perform a task other than NAND, digital organisms must make use of other instructions available, which arise through mutation or recombination. When an organism performs a task, the organism’s ‘merit’ is increased by a specific amount, specified in a configuration file, for that task. The merit of an organism is a unitless value used by Avida to determine the number of instructions an organism may execute each time step. If two organisms had the same merit, they would execute the same number of instructions at each time step; however, if one organism had twice the merit as another, the first organism would execute twice the number of instructions compared to the second in a single time step. Thus, an organism with twice the merit as another would replicate twice as fast. Organisms initially inherit the merit of their parents; otherwise, new organisms would be at a disadvantage compared to the rest of the population. The default environment rewards for nine binary (i.e., two-input) tasks. Adaptation in Avida occurs naturally (i.e., it is not simulated), as a result of the three ingredients required for natural selection: inheritance, variation, and differential reproduction. Inheritance comes from replication (sexual or asexual), variation comes from mutation and recombination, and differential reproduction comes from their rate of replication (determined by their replication code and performance of tasks). The ability to perform tasks evolves as organisms with the right mutations replicate faster than others and therefore take over the population. There are many ways in which to perform any one task, and independently evolved organisms often evolve the same task in different ways and with different degrees of efficiency. 8 Dissertation summary Evolutionary adaptation to a new environment depends on the availability of beneficial alleles. Beneficial alleles may appear as new mutations or may come from standing genetic variation—alleles already present in the population prior to the environmental change. Adaptation from standing genetic variation in sexually-reproducing populations is expected to be faster than from new mutations because beneficial alleles from standing genetic variation occur at a higher starting frequency and are immediately available. The distribution of fitness effects of alleles from standing genetic variation are expected to be different from that of new mutations because standing genetic variation has been ‘pre-tested’ by selection. Whether adaptation uses standing genetic variation or new mutations as a source of beneficial alleles is unknown. In Chapter 1 (p. 15), I conducted experimental evolution of digital organisms to determine the source of beneficial alleles during adaptation. I also tested the speed of adaptation and the fitness effect of alleles under these two sources of genetic variation. Various processes (e.g., population bottlenecks and hitchhiking) may drive a biological population to acquire mutations that reduce its fitness (i.e., “deleterious mutations”). A population with deleterious mutations may fully or partially recover in fitness in two ways: reversion or compensatory adaptation. Compensatory adaptation is the fixation of mutations that ameliorate the effects of deleterious mutations while the original deleterious mutations remain fixed. Reversion is often the best way to recover and, if a revertant mutation were to appear, the most probable route. However, it has been found experimentally that there are more compensatory mutations available than the single revertant. It has also been found that once compensatory adaptation has begun, reversion becomes an increasingly improbable route because the effect of the reversion changes with the genetic background. 9 Therefore, it seems that reversion has a limited window of opportunity in which its full effect would be beneficial. The fitness effect of the deleterious mutation, the population size, and the mutation rate are three main factors that will affect this window of opportunity. In Chapter 2 (p. 35), I used populations of digital organisms to investigate the effect of these factors. I found that the lower the initial fitness of the population, the smaller the window of opportunity for reversion. This result was partly caused by the stronger negative interactions between compensatory mutations and the reversion. I found that the window of opportunity for reversion was highest the larger the population size, but compensatory adaptation was most probable at intermediate population sizes and lowest at the extremes. Finally, I found that the higher the mutation rate, the larger the window of opportunity reversion, but it was smaller than expected because the higher mutation rate caused more negative interactions with compensatory mutations to occur. Epistatic interactions among compensatory mutations that have evolved in separate populations may form an intrinsic postzygotic isolating barrier (i.e., hybrid inviability or sterility), leading to biological speciation. In Chapter 3 (p. 66), I tested experimentally whether compensatory adaptation can lead to reproductive isolation (specifically, postzygotic isolation) and whether it was more rapid and stronger than in populations evolved through drift. Surprisingly, the strength of this isolation was independent of the effect size of the original deleterious mutations. I also find that both deleterious and compensatory mutations contribute equally to reproductive isolation. Our results suggest that compensatory adaptation may be an important genetic mechanism of speciation, and supports the view that intrinsic postzygotic isolation can stem from adaptation to the genetic environment. Reproductive isolation between populations often evolves as a byproduct of independent adaptation to new environments, but the selective pressures of these environments may be 10 divergent (‘ecological speciation’) or uniform (‘mutation-order speciation’). In Chapter 4 (p. 86), I directly compare the strength of reproductive isolation (specifically, postzygotic) generated by ecological and mutation-order processes. I also tested the effect of gene flow as well as the dimensionality (i.e., number of selective pressures) of the environments on the strength of postzygotic isolation. I found that ecological speciation generally formed stronger isolation than mutation-order speciation, mutation-order speciation was more sensitive to gene flow than ecological speciation, and environments with high dimensionality formed stronger reproductive isolation than those with low dimensionality. How various factors affect the strength of reproductive isolation has been difficult to test in biological organisms, but the use of artificial life, which provides its own genetic system that evolves, allowed us to computationally test the effect of these factors more easily. Under the Dobzhansky-Muller model of speciation, hybrid inviability or sterility results from the evolution of genetic incompatibilities (DMIs) between species-specific alleles. This model predicts that the number of pairwise DMIs between species should increase quadratically through time, but the few tests of this ‘snowball effect’ have had conflicting results. In Chapter 5 (p. 115), I show that pairwise DMIs accumulated quadratically, supporting the snowball effect. The number of unfit hybrids has been proposed to accumulate faster than quadratically because hybrids harbor more complex DMIs, but I found that the accumulation was linear. I show that more complex genetic interactions involved alleles that rescued pairwise incompatibilities, explaining the discrepancy between the accumulations of DMIs versus unfit hybrids. Our results highlight the importance of complex genetic interactions in speciation. 11 LITERATURE CITED 12 LITERATURE CITED Adami, C. (2006). Digital genetics: unraveling the genetic basis of evolution. Nature Rev. Genet., 7:109–118. Barrett, R. D. H. and Schluter, D. (2008). Adaptation from standing genetic variation. Trends Ecol. Evol., 23:38–44. Beckmann, B. E., McKinley, P. K., Knoester, D. B., and C., O. (2007). Evolution of cooperative information gathering in self-replicating digital organisms. In Proceedings of the First IEEE International Conference on Self-Adaptive and Self-Organizing Systems. Coyne, J. A. and Orr, H. A. (2004). Speciation. Sinauer Associates, Inc., Sunderland, MA. Elena, S. F. and Lenski, R. E. (2003). Evolution experiments with microorganisms: the dynamics and genetic bases of adaptation. Nature Rev. Genet., 4:457–469. Goldsby, H. J. and Cheng, B. H. C. (2008). Avida-mde: a digital evolution approach to generating models of adaptive software behavior. In GECCO’08. Knoester, D. B. and McKinley, P. K. (2011). Evolution of synchronization and desynchronization in digital organisms. Artif. Life, 17:1–20. Knoester, D. B., Ramirez, A. J., McKinley, P. K., and Chen, B. H. C. (2009). Evolution of robust data distribution among digital organisms. In GECCO’09. Lenski, R. E. (2001). Twice as natural. Nature, 414:255. McKinley, P., Cheng, B. H. C., Ofria, C., Knoester, D., Beckmann, B., and Goldsby, H. (2008). Harnessing digital evolution. Computer, 41:54–63. Misevic, D., Lenski, R. E., and Ofria, C. (2004). Sexual reproduction and mullers ratchet in digital organisms. In Artificial Life IX. Misevic, D., Ofria, C., and Lenski, R. E. (2006). Sexual reproduction reshapes the genetic architecture of digital organisms. Proc. R. Soc. B, 273:457–464. Misevic, D., Ofria, C., and Lenski, R. E. (2010). Experiments with digital organisms on the origin and maintenance of sex in changing environments. J. Hered., 101:S46–S54. Ofria, C. and Wilke, C. O. (2004). Avida: a software platform for research in computational evolutionary biology. Artif. Life, 10:191–229. O’Neill, B. (2003). Digital evolution. PLoS Biol., 1:011–014. Orr, H. A. (2005). The genetic theory of adaptation: a brief history. Nature Rev. Genet., 6:119–127. Pennock, R. T. (2007). Models, simulations, instantiations, and evidence: the case of digital evolution. J. Exp. Theor. Artif. Intell., 19:29–42. 13 Schemske, D. W. (2010). Adaptation and The Origin of Species. Amer. Nat., 176:S4–S25. Schluter, D. (2009). Evidence for ecological speciation and its alternative. Science, 323:737– 741. Sobel, J. M., Chen, G. F., Watt, L. R., and Schemske, D. W. (2010). The biology of speciation. Evolution, 64:295–315. Wilke, C. O. and Adami, C. (2002). The biology of digital organisms. Trends Ecol. Evol., 17:528–532. Yedid, G. and Bell, G. (2001). Microevolution in an electronic microcosm. Am. Nat., 157:465–487. 14 Chapter 1 The role of standing genetic variation in adaptation to a new environment 1.1 Introduction When a population adapts to a new environment, beneficial alleles may appear as new mutations or come from standing genetic variation (Barrett and Schluter, 2008). Standing genetic variation refers to the presence of alternative alleles at each genetic locus in a population. Standing genetic variation may be maintained in a population for several reasons (Hartl and Clark, 1997); e.g., alleles with little or no effect on fitness may rise to moderate frequencies by random genetic drift. Standing genetic variation may be a major source of beneficial alleles in a new environment, with two important implications for the dynamics of adaptation. First, adaptation from standing genetic variation should be faster than adaptation from new mutations because beneficial alleles would be immediately available and would be present at higher frequencies (Barrett and Schluter, 2008). Second, the distribution of fitness effects of alleles from standing genetic variation should be different than that of new mutations because standing genetic variation has been ‘pre-tested’ by surviving previous generations of selection against deleterious alleles (Barrett and Schluter, 2008). 15 Whether standing genetic variation is an important source of beneficial alleles for adaptation is unknown. Studies have employed three main approaches to answer this question (reviewed in Barrett and Schluter (2008)): analysis of the signature of selection, presence of the beneficial allele in the ancestral population, and phylogenetic analysis for inferring the history of alleles. These methods, however, are necessarily indirect and each has their unique set of problems. Of course, the “surest way to determine the source of beneficial alleles is to locate the genes themselves and establish their histories” (Barrett and Schluter, 2008). In this study, I used digital organisms (see p. 6) to follow individual alleles through time as populations adapted to a new environment, and I determined whether beneficial alleles appeared as new mutations or came from standing genetic variation. I also tested whether adaptation from standing genetic variation was faster than from new mutations and whether the fitness effects of standing genetic variation were different from those of new mutations. 1.2 Standing genetic variation in digital organisms To generate a well-adapted, sexual population with standing genetic variation prior to the environmental change, I initialized an empty ‘world’ with an organism that could replicate but could not perform any tasks. I set the world size to 10,000 cells and the environment to reward for the default nine tasks (Lenski et al., 1999). I set the copy mutation rate to 0.1 mutations per genome per generation and, to ensure homologous recombination, I fixed the length of all genomes to 200 instructions and turned off insertion and deletion mutations. I let 50 such replicate populations evolve for 500,000 updates—a measurement of time in Avida—which was about 42,000 generations. I then picked a random population in which the consensus sequence could perform all nine tasks (35 out of the 50 could perform all nine 16 tasks), and I took a random sample of 1,000 individuals from this population to serve as the ancestral population before the environmental change. To measure the amount of standing genetic variation in the ancestral population, I measured the heterozygosity of each locus of the population. The heterozygosity of a locus is H = 1− k 2 i=1 pi , where k is the number of alleles segregating at that locus and pi is the frequency of the ith allele (Gillespie, 2004, p. 15). Here I adopted the convention that a locus is polymorphic (i.e., has standing genetic variation) if its most common allele has a frequency < 0.95 (Hartl and Clark, 1997, p. 53). A locus that had standing genetic variation would have a minimum heterozygosity of 1 − (0.952 + 0.052 ) = 0.095. Because there are 26 possible alleles (i.e., instructions) per locus in digital organisms, the maximum possible heterozygosity is approximately 0.9615. I found substantial standing genetic variation in the ancestral population (Figure 1.1). Of 200 loci, 125 (62.5%) were polymorphic. The heterozygosity of each locus ranged from 0.0 to 0.8859, with a mean heterozygosity of 0.3781 (0.3334–0.4246, 95% bootstrap CI). For comparison, Stephens et al. (2001) found in humans that the heterozygosity of 313 genes ranged from 0.012 to 0.929, with a mean of 0.534. In natural populations of E. coli, Selander and Levin (1980) found that the heterozygosity of 20 enzyme-encoding genes ranged from 0.055 to 0.887, with a mean of 0.4718. My results demonstrate that the ancestral population exhibited levels of standing genetic variation consistent with that observed in biological populations. Furthermore, they support the claim that standing genetic variation is a ubiquitous property of evolving genetic systems (Gibson and Dworkin, 2004; Barrett and Schluter, 2008). 17 1.3 Source of beneficial alleles Having established that the ancestral population harbored abundant standing genetic variation, I determined whether adaptation to a new environment relied on this genetic variation or on new mutations as a source of beneficial alleles. In this study, I examined beneficial alleles with fitness effects greater than 1%. With the ancestral population, I started 20 new replicate populations in a world of 1,000 cells and an environment that rewarded for 68 different tasks (the original nine tasks were not rewarded for). As a control, I also started another set of 20 replicate populations where every individual had an identical genotype (i.e., isogenic), set to the consensus sequence of the ancestral population. Although the consensus genotype did not actually exist in the ancestral population, its fitness was 1.0070 relative to the highest fit individual in the ancestral population (excluding those who could immediately perform tasks), and 1.0337 relative to the mean fitness of the ancestral population. Thus, the control population was not at a disadvantage compared to the ancestral population. All other configuration settings were identical to those used for the evolution of the ancestral population. Note that the populations that started with standing genetic variation were also Heterozygosity allowed to get new mutations (the mutation rate was set to 0.1 mutations per genome per 1.0 0.5 0.0 1 20 40 60 80 100 Locus 120 140 160 180 200 Figure 1.1: The heterozygosity of each locus of the population before the environmental change. Heterozygosities above 0.095 indicate the presence of standing genetic variation. 18 generation). I let these replicate populations evolve for 10,000 updates (∼ 850 generations), saving each population every 100 updates. At the end of the runs, I found that the populations that started with standing genetic variation increased in mean fitness to 8.31 (7.74–8.87, 95% bootstrap CI) relative to the ancestral population in the new environment (i.e., the evolved populations were 8.31 times more fit in the new environment than the ancestral population). These populations were able to perform an average of 7.9 tasks, with a range of 5 to 10. The mean number of fixed, derived alleles—defined as having a frequency > 0.95 in the evolved population but < 0.95 in the ancestral population—was 56.25, ranging from 38 to 70. Figure 1.2 shows the history of two allele fixation events, one from standing genetic variation and the other from a new mutation, that occurred in the first replicate population. Of the 56.25 fixed, derived alleles, 47.8 (85%) existed as standing genetic variation in the ancestral population. In the control populations, mean fitness increased to 7.18 (6.62–7.76, 95% bootstrap CI) relative to the ancestral population. The control populations were able to perform an average of 6.7 tasks, with a range of 5 to 9. The mean number of fixed, derived alleles in the control populations was 5.15, ranging from 2 to 9. It was surprising that the populations that started with standing genetic variation fixed 10 times more alleles than the control populations, despite both sets of populations having similar final fitnesses and number of tasks performed. The finding that 85% of fixed, derived alleles in the populations that started with the ancestral population existed as standing genetic variation may indicate that most beneficial alleles came from standing genetic variation. It is not clear, however, whether they were fixed by neutral genetic drift, natural selection, or genetic linkage and hitchhiking with beneficial alleles. For example, genetic hitchhiking in Avida can occur when alleles nearby a highly beneficial allele rise in frequency along with the beneficial allele. Hitchhiking occurs 19 Frequency 0 2,500 5,000 Time (updates) 7,500 10,000 Figure 1.2: The frequencies of alleles through time for two loci in which an allele became beneficial and subsequently fixed. In the top plot, the beneficial allele came from standing genetic variation, and in the bottom plot, the beneficial allele appeared as a new mutation. Different alleles are represented by different colors. The y-axis in each plot ranges from 0.0 to 1.0. For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation. because the beneficial allele and nearby (i.e., genetically linked) alleles spread faster than recombination can break them apart. It is also not clear at what frequency the derived alleles first became beneficial. Therefore, I developed a method to systematically measure the fitness of individual alleles through time and determine the frequency at which they became beneficial. First, for each fixed, derived allele at the end of each run, I calculated both the allele’s frequency and fitness effect every 100 updates, starting at the first update. To calculate the fitness effect of an allele at the current update, I first selected from the population the individual with the highest fitness who had the allele. I then created a clone of the individual and substituted the allele with an alternative allele drawn randomly from the standing genetic variation at that locus. I then calculated the fitness of the individual with the allele relative to the fitness of the individual without it. If this relative fitness was greater than 1.01, then the fitness effect of the allele (> 1%) was beneficial at the current update. While testing this method, I found some cases where the fitness effect of the allele was considered beneficial only because the individual with the alternative allele had unusually low fitness. To reduce 20 Cumulative frequency of occurrence 1.0 0.8 0.773 0.6 0.4 0.2 0.134 0.0 Populations started with SGV Populations started without SGV 0.0 1.0 0.2 0.4 0.6 0.8 0.05 Frequency of allele when it became beneficial Figure 1.3: The cumulative frequency of fixed alleles that became beneficial at a specific frequency (0.05 bin size) for populations that started with standing genetic variation (solid lines) and for control, isogenic populations (dashed lines). The gray lines indicate the 95% bootstrap confidence interval around the mean of 20 replicate populations. The red vertical line indicates the frequency below which alleles were considered to appear as new mutations. The red horizontal lines indicate the proportions of alleles that came from new mutations for either type of population. the frequency of such cases, I also required that the allele be beneficial for the individual with the second highest fitness. I stopped analyzing further updates as soon as I found the allele to be beneficial or if it became fixed. In populations that started with standing genetic variation, I found that out of the mean 56.25 alleles that fixed, a mean of 31.9 became beneficial at some point in their history. I found that only 13.4% of these beneficial alleles became beneficial at a frequency < 0.05 (Figure 1.3, lower horizontal red line); the remaining 86.6% became beneficial at a frequency > 0.05. Supposing standing genetic variation comprises alleles with frequencies > 0.05, these results indicate that the majority of beneficial alleles came from standing genetic variation. In the control populations, I found that out of the mean 5.15 alleles that fixed, a mean of 5.1 21 became beneficial at some point in their history. I found that 77.3% of these beneficial alleles became beneficial at a frequency < 0.05 (Figure 1.3, upper horizontal red line); the remaining 22.7% became beneficial at a frequency > 0.05. Therefore, in contrast to populations that started with standing genetic variation, the control, isogenic populations adapted mostly from new mutations, although almost a quarter of beneficial alleles came from standing genetic variation that arose as populations accumulated genetic polymorphism over time. Interestingly, the mean absolute (not percentage) number of new mutations per replicate for each treatment was about the same: 4.15 (3.40–4.85, 95% bootstrap CI) for populations started with standing genetic variation and 3.75 (3.3–4.2) for isogenic populations. This indicates that standing genetic variation did not inhibit new mutations from being selected. One potential concern with the above method is that I identified beneficial alleles based on only two genotypes that had the allele, relative to two genotypes with alternative alleles. Yet the presumed beneficial alleles as well as the alternative alleles may not have the same fitness effect on other genetic backgrounds. Thus, I implemented a second method to identify beneficial alleles that considered more genotypes when measuring fitness effects. The key difference between this method and the previous is that in this method I selected all individuals who had the allele. Then, for each of these individuals I substituted the allele with an alternative allele drawn randomly from the standing genetic variation at that locus. Finally, I calculated the mean fitness of all individuals with the allele relative to the mean fitness of all individuals with the allele replaced. If this relative fitness was greater than 1.01, then I considered the allele as beneficial. Using this method, I found that in populations that started with standing genetic variation, 11.5% of alleles became beneficial at a frequency < 0.05; the remaining 88.5% became beneficial at a frequency > 0.05. In the isogenic populations, I found that 79.4% of alleles became beneficial at a frequency < 0.05; 22 the remaining 20.6% became beneficial at a frequency > 0.05. These results are very similar to those I found with the previous method, showing that the previous method was robust to the number of genotypes considered when identifying beneficial alleles. 1.4 Speed of adaptation Adaptation from standing genetic variation should be faster than adaptation from new mutations because beneficial alleles would be immediately available and would be present at higher frequencies (Barrett and Schluter, 2008). To test this prediction, I compared the speed of adaptation between populations that started with standing genetic variation and those that started with isogenic individuals. I re-evolved both types of populations at the additional mutation rates (U ) of 0.01 and 0.0 (no new mutations) per genome per generation (the original populations were run at a mutation rate of 0.1). I added these new treatments because, given that the only source of mutations for the isogenic populations were new mutations, the mutation rate would be an important variable on the rate of adaptation. Population size would also be an important variable on the rate of adaptation, but I did not investigate its effects in this study. I found that at the 0.1 mutation rate, the rate of adaptation for populations that started with standing genetic variation was significantly greater for most of the first four thousand updates than isogenic populations, then became less significantly so for the rest of the run (Figure 1.4A). At the 0.01 mutation rate, however, the rate of adaptation was significantly greater for the entire run (Figure 1.4B). Interestingly, at the 0.0 mutation rate, populations with standing genetic variation continued to adapt for several thousand updates, but, as expected, isogenic populations could not evolve (Figure 1.4C). These results clearly demon- 23 strate that adaptation from standing genetic variation was faster than from new mutations. Yet new mutations were necessary for long-term evolution, as shown by the fact that adaptation from standing genetic variation without new mutations stopped after several thousand updates. 1.5 Fitness effect of random alleles from different sources of variation The distribution of fitness effects of alleles from standing genetic variation should be different than that of new mutations because standing genetic variation has been ‘pre-tested’ by selection (Barrett and Schluter, 2008). To test this prediction, I generated the fitness effect distribution of alleles coming from either standing genetic variation or new mutations, measured in the new environment. First, I sampled 1,000 random (but viable) individuals from the ancestral population and mutated a single, random locus of each individual to an allele drawn randomly from the standing genetic variation (if there was any variation at that locus). I also sampled another set of 1,000 individuals from the ancestral population and mutated a single locus of each individual to an allele drawn randomly from all 25 possible alternative alleles. To prevent the possibility that these random mutations were more deleterious only because they disrupted fixed alleles, I ensured that the loci were drawn from the same pool of loci that had standing genetic variation. Finally, I measured the fitness of these mutants relative to the original, unmutated individual. I found that the mean fitness of mutants with mutations from standing genetic variation was 0.9994 (0.9969–1.0023, 95% bootstrap CI). The mean fitness of mutants with random mutations was 0.9496 (0.9326–0.9665, 95% bootstrap CI). Clearly, mutations from standing 24 10 8 A U = 0.1 6 4 2 Mean relative fitness 0 10 8 5,000 7,500 10,000 2,500 5,000 7,500 10,000 2,500 5,000 7,500 10,000 2,500 B U = 0.01 6 4 2 0 10 8 C U = 0.0 6 4 2 0 Time (updates) Normal SGV No SGV Figure 1.4: The mean fitnesses (relative to the ancestor) of populations evolved after an environmental change at (A) 0.1, (B) 0.01, and (C) 0.0 mutations per genome per generation (U ). Populations evolved starting either with the ancestral population (solid line), which contained standing genetic variation (SGV) or with an isogenic population based on the consensus sequence of the ancestral population (dashed line). Gray lines represent the 95% bootstrap confidence intervals around the mean. 25 Fitness effect Lethal Strongly deleterious Mildly deleterious Nearly neutral Mildly beneficial Strongly beneficial Source of mutation SGV Random 0 58 3 5 186 345 729 520 81 67 1 5 Table 1.1: The number of single mutants (out of 1,000), categorized by the mutation’s source and fitness effect (w): lethal (w = 0), strongly deleterious (0 < w ≤ 0.99), mildly deleterious (0.99 < w ≤ 0.999), neutral or nearly neutral (0.999 < w ≤ 1.001), mildly beneficial (1.001 < w ≤ 1.01), and strongly beneficial (w > 1.01). genetic variation did not have, on average, as strong deleterious effects as random mutations. To examine more closely the fitness effects of mutations from the two sources, I categorized each mutation based on the mutant’s relative fitness (Table 1.1). Alleles from standing genetic variation were mostly neutral, whereas new mutations were more likely to be lethal or deleterious. Interestingly, new mutations were also more likely to be strongly beneficial than alleles from standing genetic variation, yet in the analysis where I determined the source of beneficial alleles, I found that most beneficial alleles came from standing genetic variation. This discrepancy may indicate that although alleles from standing genetic variation were not beneficial alone, combinations of these alleles brought together by recombination provided the benefits. The finding that alleles from standing genetic variation were less deleterious on average than random mutations support the hypothesis that standing genetic variation has been pre-tested by selection. The above analysis was based on randomly generated mutants of the ancestral genotypes (i.e., at the beginning of the experiments), but it would also be interesting to know the fitness effect of beneficial alleles that actually fixed. This information was already calculated as part of determining the moment at which alleles became beneficial because it was used to deter26 mine whether alleles had achieved a fitness > 1.01 (using the first method). For populations that had evolved under standing genetic variation, the mean fitness of a genotype with a beneficial allele at the moment at which it became beneficial (relative to a genotype without the beneficial allele) was 1.54 (1.48–1.60, 95% bootstrap CI). For isogenic populations, this mean fitness was 1.47 (1.37–1.59, 95% bootstrap CI). Although the mean fitness effect of beneficial alleles for the standing genetic variation treatment was slightly higher than the isogenic treatment, they were not significantly different. The maximum relative fitness for a genotype with a beneficial allele for the standing genetic variation treatment (7.05) was higher than that for the isogenic treatment (4.50). 1.6 Discussion I have shown that in populations of digital organisms adapting to a new environment, the major source of beneficial alleles was standing genetic variation, not new mutations. My findings are supported by selection experiments and observational studies of biological populations. Selection experiments have shown that adaptation can occur by changes in allele frequencies of standing genetic variation in the initial populations (e.g., Feder et al., 1997; Scarcelli and Kover, 2009; Teot´nio et al., 2009). Observational studies of natural populao tions have found that alleles correlated with adaptive traits were also present in the ancestral population (e.g., Colosimo et al., 2005; Myles et al., 2005). In biological organisms, however, it is very difficult to measure the fitness effects of individual alleles, which is necessary to determine whether an allele fixed due to selection. Another problem, specific to studies of natural populations, is that the ancestral population is unavailable—the closest one can get is the extant population from which a subpopulation founded a new environment—and 27 therefore it is often unknown whether a beneficial allele existed as standing genetic variation. The use of digital organisms allowed me to track individual alleles through time and determine the frequency at which they became beneficial. When alleles from standing genetic variation became beneficial, their starting frequency ranged from the minimum of 5% to the maximum of 95% (Figure 1.3). In experimental studies of biological organisms, high starting frequencies (> 50%) are not uncommon (e.g., Feder et al., 1997; Scarcelli and Kover, 2009). In natural populations, however, starting frequencies have tended to be much smaller, such as in the study by Colosimo et al. (2005), where the starting frequency of an adaptive allele was between 0.2% and 3.8% in the ancestral population. One possible reason for this discrepancy is that natural populations may be under stronger selective pressures than experimental populations (Ellegren and Sheldon, 2008), so the fitness effects of alleles in natural populations tend to be more deleterious and therefore maintained at low frequencies. Of course, allele frequency data for adaptive alleles in natural populations is scarce, so more research in natural populations should determine the frequencies at which alleles from standing genetic variation become beneficial. Adaptation should be faster if most beneficial alleles came from standing genetic variation than if they came from new mutations (Barrett and Schluter, 2008). I found this to be the case in digital organisms if the mutation rate was low enough (Figure 1.4). In fact, when no new mutations were allowed, adaptation by standing genetic variation continued for several hundred generations, whereas no adaptation occurred in isogenic populations. Still, the importance of new mutations for long-term evolution was shown by the fact that adaptation stopped eventually when no new mutations were allowed. Although there are no empirical studies testing the speeds of adaptation, where beneficial alleles may come from either standing genetic variation or new mutations, my results are supported theoretically (Hermisson 28 and Pennings, 2005). There are two reasons that adaptation from standing genetic variation should be faster than adaptation from new mutations: beneficial alleles are both readily available and present at higher frequencies than alleles from new mutations (Barrett and Schluter, 2008), which must overcome drift because they start at lower frequencies. Future experiments should be able to quantify the relative contribution of these two causes. Although not examined in detail in this study, the population size and mutation rate can affect the relative contributions of standing genetic variation and new mutations during adaptation. For example, a sudden decrease in population size (i.e., a bottleneck) will reduce both the amount of standing genetic variation and the number of new mutations that appear each generation. In this case, standing genetic variation will still have an advantage over new mutations—especially for alleles of weak fitness effect—because weak effect alleles introduced by new mutations are easily lost due to genetic drift (Hermisson and Pennings, 2005). For large effect alleles, standing genetic variation will have a reduced advantage because large effect alleles are less likely to be lost even if they are introduced as new mutations (Hermisson and Pennings, 2005). In my experiments, mutations that allowed organisms to perform new tasks were of large effect (the default configuration in Avida), but future studies should experiment with weaker beneficial alleles. In a large population or high mutation rate, new mutations would become more important because large-effect mutations would appear more frequently. Because alleles from standing genetic variation have had a potentially long history in an evolving population, their fitness effects in a new environment have been predicted to be less deleterious than random mutations (Barrett and Schluter, 2008). On average, I found that standing genetic variation was effectively neutral (fitness effect of 0.0006), whereas random mutations were strongly deleterious (fitness effect of 0.0504). Alleles from standing genetic 29 variation can therefore linger in a population, increasing the chance for them to become beneficial after an environmental or genetic change. Random mutations, on the other hand, are on average deleterious and are thus more easily eliminated by selection. In biological populations, the mean fitness effect of random mutations was found to be 0.48 in RNA viruses (Sanju´n et al., 2004), 0.12 in C. elegans (Vassilieva et al., 2000), and 0.22 in yeast a (Zeyl and DeVisser, 2001). There are no measurements of the fitness effects of alleles from standing genetic variation in a biological population in a new environment. For strongly beneficial mutations (i.e., fitness effect > 1%), I found that random mutations were more likely to be beneficial than alleles from standing genetic variation in the new environment (Table 1.1). It may thus seem counter-intuitive that most beneficial alleles during adaptation came from standing genetic variation. I hypothesize that it was the combination of many alleles from standing genetic variation that provided the benefits, and together these epistatically related alleles rose to fixation. Adaptation that requires many alleles working together is known as ‘polygenic adaptation’ (Pritchard and Di Rienzo, 2010), although fixation of alleles is not always necessary. In fact, Pritchard and Di Rienzo (2010) hypothesize that if adaptation occurs from standing genetic variation, polygenic adaptation is likely. In summary, this study has shown the importance of standing genetic variation in populations of digital organisms adapting to a new environment. That is, (1) most beneficial alleles came from standing genetic variation rather than from new mutations, (2) populations that started with standing genetic variation adapted faster than populations that started with identical genotypes, and (3) the fitness effects of alleles from standing genetic variation were less harmful than new mutations. Because digital organisms evolve by the same processes of natural selection and genetic drift that biological populations also experience, I suspect 30 that the above points are also true for biological populations. A hypothesis that arose from this study was that standing genetic variation together with recombination may give rise to combinations of alleles that together are beneficial. Future work should test whether this additional advantage is true, thereby highlighting the importance of sexual recombination and standing genetic variation in evolving populations. 31 LITERATURE CITED 32 LITERATURE CITED Barrett, R. D. H. and Schluter, D. (2008). Adaptation from standing genetic variation. Trends Ecol. Evol., 23:38–44. Colosimo, P. F., Hosemann, K. E., Balabhadra, S., Villarreal Jr., G., Dickson, M., Grimwood, J., Schmutz, J., Myers, R. M., Schluter, D., and Kingsley, D. M. (2005). Widespread parallel evolution in sticklebacks by repeated fixation of ectodysplasin alleles. Science, 307:1928–1933. Ellegren, H. and Sheldon, B. C. (2008). Genetic basis of fitness differences in natural populations. Nature, 452:169–175. Feder, J. L., Roethele, J. B., Wlazlo, B., and Berlocher, S. H. (1997). Selective maintenance of allozygme differences among sympatric host races of the apple maggot fly. Proc. Natl. Acad. Sci. USA, 94:11417–11421. Gibson, G. and Dworkin, I. (2004). Uncovering cryptic genetic variation. Nature Rev. Genet., 4:681–690. Gillespie, J. H. (2004). Population genetics: a concise guide. Johns Hopkins University Press, Baltimore, MD. Hartl, D. L. and Clark, A. G. (1997). Principles of population genetics. Sinauer Associaties, Inc., Sunderland, MA. Hermisson, J. and Pennings, P. S. (2005). Soft sweeps: molecular population genetics of adaptation from standing genetic variation. Genetics, 169:2335–2352. Lenski, R. E., Ofria, C., Collier, T. C., and Adami, C. (1999). Genome complexity, robustness, and genetic interactions in digital organisms. Nature, 400:661–664. Myles, S., Bouzekri, N., Haverfield, E., Cherkaoui, M., Dugoujon, J.-M., and Ward, R. (2005). Genetic evidence in support of a shared eurasian-north african dairying origin. Hum. Genet., 117:34–42. Pritchard, J. K. and Di Rienzo, A. (2010). Adaptation – not by sweeps alone. Nature Rev. Genet., 11:665–667. Sanju´n, R., Moya, A., and Elena, S. F. (2004). The distribution of fitness effects caused by a single-nucleotide substitutions in an RNA virus. Proc. Natl. Acad. Sci. USA, 101:8396– 8401. Scarcelli, N. and Kover, P. X. (2009). Standing genetic variation in FRIGIDA mediates experimental evolution of flowering time in Arabidopsis. Mol. Ecol., 18:2039–2049. Selander, R. K. and Levin, B. R. (1980). Genetic diversity and structure in Escherichia coli populations. Science, 210:545–547. 33 Stephens, J. C. et al. (2001). Haplotype variation and linkage disequilibrium in 313 human genes. Science, 293:489–493. Teot´nio, H., Chelo, I. M., Bradi´, M., Rose, M. R., and Long, A. D. (2009). Experimental o c evolution reveals natural selection on standing genetic variation. Nat. Genet., 41:251–257. Vassilieva, L. L., Hook, A. M., and Lynch, M. (2000). The fitness effects of spontaneous mutations in Caenorhabditis elegans. Evolution, 54:1234–1246. Zeyl, C. and DeVisser, J. A. G. M. (2001). Estimates of the rate and distribution of fitness effects of spontaneous mutation in Saccharomyces cerevisiae. Genetics, 157:53–61. 34 Chapter 2 The probability of compensatory adaptation in digital organisms 2.1 Introduction Deleterious mutations may accumulate in a population through various mechanisms: genetic drift (Lande, 1994; Lynch et al., 1995), hitchhiking with beneficial mutations (Chun and Fay, 2011), transient environmental changes (Bj¨rkman et al., 2000), and selfish genetic elements o (Presgraves, 2010). In these deteriorated populations, new beneficial mutations may arise and restore the fitness of the population. These beneficial mutations, however, may not have been beneficial in the absence of the accumulated deleterious mutations. In other words, their effect may epistatically depend on the current genetic background, and without the deleterious mutations, they may have had no benefit or even have been deleterious. These beneficial mutations are known as ‘compensatory mutations,’ and the process by which compensatory mutations recover fitness is known as ‘compensatory adaptation.’ Experimental evolution studies have observed compensatory adaptation (Hartl and Taubes, 1996; Burch and Chao, 1999; Moore et al., 2000; Levin et al., 2000; Maisnier-Patin et al., 2002; Estes and Lynch, 2003; Estes et al., 2011). 35 Understanding compensatory adaptation has practical applications to society, such as the conservation of threatened or endangered species and the antibiotic resistance of pathogenic bacteria. Theoretical work has shown that small populations (< 100 individuals) or populations that have undergone bottlenecks will readily fix deleterious mutations (Whitlock et al., 2003). There is the risk that such populations will go extinct (Lynch et al., 1995; Lande, 1994), unless compensatory adaptation could help them recover. Bacteria susceptible to an antibiotic may acquire resistance mutations in the presence of the antibiotic, but such mutations are often deleterious in the absence of the antibiotic (Schrag et al., 1997; Levin et al., 2000). The hope for combating resistant bacteria was to remove the antibiotic, so that a competitively superior susceptible strain would evolve through reversion. Resistant bacteria, however, may instead acquire compensatory mutations that remove this fitness deficit while retaining their resistance (Schrag et al., 1997; Levin et al., 2000; Bj¨rkman o et al., 2000; Maisnier-Patin et al., 2002; Paulander et al., 2007; Perron et al., 2010). In fact, because compensatory mutations often depend on the deleterious mutations already present, reversion and susceptibility become increasingly difficult. Experimental studies have found that compensation, rather than reversion, often occurs, but reversion is sometimes present or inferred (Burch and Chao, 1999; Sanju´n et al., 2005; a Maisnier-Patin et al., 2002). The reason that reversion is rare has been argued to be that compensatory mutations are much more frequent than revertant mutations (i.e., there is only one way to revert, but many ways to compensate) (Levin et al., 2000; Whitlock et al., 2003; Sanju´n et al., 2005). One might expect, however, that once the revertant mutation appears a it would most likely fix because its fitness recovery is 100% (i.e., its selective value would be higher than that of compensation). However, epistasis is common in organisms, and the possibility exists that once a few compensatory mutations arise and fix, the revertant muta36 tion will not provide its full benefits because it would interact negatively with compensatory mutations. This possibility has been observed experimentally (Schrag et al., 1997; Levin et al., 2000). The conflict between compensatory and reversion is a complicated interaction involving, at least, the initial fitness of the mutant, population size, and mutation rate. Here I examine the effect of these factors one by one on the probability of compensation vs. reversion. The initial fitness is important because, primarily, it will determine the fitness effect of the reversion when it appears, assuming this fitness has not changed because of compensatory mutations. It is also important because the number of compensatory mutations may be different, given that it is expected that there will be more ways to compensate the lower the fitness of the mutant. The population size is important because it partly determines how many mutations arise each generation. Larger populations will increase both the frequency of compensatory mutations and that of reversion, and studies have confirmed that reversion occurs more frequently in large populations (Burch and Chao, 1999). Like population size, mutation rate also affects the frequency in which mutations arise. However, greater mutation rates increase the chance that double-mutants appear, such that a revertant mutation may arise with a deleterious mutation and thus cancel out its benefit. In this study, I used experimental evolution in silico using the artificial life system Avida (Ofria and Wilke, 2004) (see p. 6) to examine the evolutionary dynamics of compensatory adaptation. The use of digital organisms allowed us to answer questions that would be difficult even with microbial organisms. With digital organisms, one can observe hundreds of generations in a few minutes, conduct hundreds of replicate experiments, easily manipulate genomes, and accurately measure fitness. Although the system I used is artificial, it has been shown that several biological phenomena emerge naturally in Avida (Wilke and Adami, 2002; 37 Adami, 2006). Digital organisms improve on mathematical models of adaptation because in this system traits are complex, involving multiple loci and epistatic interactions among alleles (Lenski et al., 1999). In these experiments, populations of mutants with a single deleterious mutation were allowed to evolve for ∼ 850 generations. I first estimated the availability of compensatory mutations depending on the initial fitness of the mutant. Then, I examined the effect of three variables—the initial fitness of the mutant, the population size, and the mutation rate—on the probability of compensation vs. reversion. From these experiments, I hope to learn the conditions under which compensation or reversion are likely to occur. I hope to understand whether compensatory mutations could change the window of opportunity for reversion, either because compensatory mutations arise much more frequently or because negative epistasis is common. 2.2 Results I evolved two ancestral populations of digital organisms (see Methods), each composed of 10,000 individuals, to the default environment for 500,000 updates (∼ 40,000 generations). Each individual was haploid, reproduced asexually, had a genome length of 200, and had a mutation rate of 0.1 per genome per generation. Both ancestral populations had evolved long enough to fully adapt and population fitness had stabilized for thousands of generations; therefore, I consider each of these populations to be near their respective optimal fitness for this environment. Although at the phenotypic level both populations had evolved to similar fitnesses, at the genomic level, they exhibited only 13.5% identity, meaning that they represent independent experimental organisms. The consensus sequence of these two 38 ancestors served as the ancestral genotypes for subsequent experiments. For each of these ancestral genotypes, I identified five sets of two mutants, each pair with approximately a fitness of either 0.1, 0.3, 0.5, 0.7, and 0.9 relative to the ancestor, for a total of 20 mutant genotypes. 2.2.1 Frequency of compensatory mutations Before I started the experimental evolution of mutants, I confirmed in digital organisms theoretical and empirical expectations about the availability of compensatory mutations in other organisms: there should be more compensatory mutations the lower the initial fitness (i.e., there are more ways to compensate the lower your fitness). This expectation has been made theoretically from extreme value theory, where the tail of any distribution (in this case, the distribution of beneficial mutations), follows the same extreme value distribution (Orr, 2002), so the higher the fitness, the fewer beneficial mutations that are found. Empirical studies (e.g., Elena et al., 1998; Burch and Chao, 1999) have also shown that populations with deleterious mutations readily gain more beneficial mutations than fit populations (Whitlock, 2000). To test this hypothesis, I introduced every possible single mutation on each mutant and counted the number that were fully or partially compensatory (i.e., fitness increased in the presence of the deleterious mutation). I confirmed that in digital organisms, as in biological, there were more compensatory mutations available in the mutants with low fitness than in those with high fitness (Figure 2.1). These results also confirm that there are more compensatory mutations than reversion (which is always exactly one), corroborating suggestions that compensation is more likely than reversion (Maisnier-Patin et al., 2002). 39 0.12 P < 0.01, R 2 = 0.96 P = 0.07, R 2 = 0.72 Proportion of one-step mutations that are compensatory 0.10 0.08 0.06 0.04 0.02 0.00 A 0.1 0.3 0.12 0.5 0.7 0.9 P = 0.05, R 2 = 0.76 P < 0.01, R 2 = 0.92 0.10 0.08 0.06 0.04 0.02 0.00 B 0.1 0.3 0.5 0.7 0.9 Initial fitness of mutant (relative to ancestor) Figure 2.1: Proportion of one-step mutations that are compensatory. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are the 95% confidence interval of the mean. 40 2.2.2 Effect of initial fitness of mutant I evolved these 20 mutant genotypes independently for 10,000 updates (∼ 850 generations) under the same environment as the original ancestors. For each genotype, the starting population was composed of 1,000 genetically-identical individuals, and the experimental evolution for each genotype was replicated 100 times. I found that the probability of compensatory adaptation declined with the fitness of the initial mutant (Figures 2.2A and 2.2B). In contrast, the probability of reversion increased with the fitness of the initial mutant (Figures 2.2C and 2.2D). Note that because in some runs the fitness values decreased or did not change, the addition of compensatory and reversion probabilities did not always add up to 100. Note also that there is a difference in the probability of compensation for W = 0.9 between the first and second ancestor. The reason for this difference appears to be that for the first ancestor, neither compensation nor reversion occur—the population does not change fitness. This is may be due to the fact that there are few compensatory mutations available for ancestor 1 for W = 0.9 (see Figure 2.1). There are several reasons that would explain why compensation is higher than reversion the lower the fitness of the original mutant. It is important to note that the probability that a reversion fixes is determined by its selective advantage, which itself depends on the fitness of the population and on whether there is negative epistasis with current mutations. The fitness of the population is initially set by the treatment but it changes during the experiment depending on the speed of compensation. The speed of compensation is directly affected by the initial fitness of the population: the lower the initial fitness the faster the population compensates and thus increases in fitness. Negative epistasis is an intrinsic property of mutations, but the total amount of negative epistasis is affected by the number of mutations 41 Percent of populations that compensated Ancestor 1 100 100 80 80 60 60 40 40 20 20 0 A Percent of populations that reverted 0.1 100 0 0.3 0.5 0.7 0.9 B 0.3 0.5 0.7 0.9 0.1 0.3 0.5 0.7 0.9 0.5 0.7 0.9 0.1 0.3 0.5 Initial fitness of population (relative to ancestor) 0.7 0.9 100 C 80 60 60 40 40 20 20 0 D 80 0 0.1 Final relative fitness of compensated populations Ancestor 2 1.0 0.3 0.5 0.7 0.9 0.1 1.0 E 0.9 0.8 0.8 0.7 0.7 0.6 0.6 0.5 F 0.9 0.5 0.1 0.3 Figure 2.2: Effect of initial fitness on compensation. Top figures: The proportion of runs that fully or partially compensated for (A) ancestor 1 (line is best fit linear model, P = 0.002) and (B) ancestor 2 (P = 0.043). Middle figures: The proportion of runs that reverted for (C) ancestor 1 (P < 0.001) and (D) ancestor 2 (P = 0.072). Bottom figures: The proportional increase in fitness for compensatory runs for (E) ancestor 1 and (F) ancestor 2. 42 present when the reversion arrives. This number of mutations depends on the speed of compensation and the rate of compensatory mutations, both of which are partly determined by the initial fitness of the population. Therefore, there are both direct and indirect ways in which the lower the initial fitness of the population the lower the probability that a reversion fixes. If negative epistasis was present, the expectation is that reversion would stop being beneficial as mutations fixed in the population. The sooner reversions stopped being beneficial the stronger that negative epistasis is with other mutations. To test whether negative epistasis may have contributed in preventing reversions from spreading in the evolving populations, I calculated the number of generations in which a reversion stopped being beneficial in populations that did not eventually revert. First, I reverted the initial mutation to the ancestral state for every individual in the population at each update saved (every 100 updates or ∼ 8.5 generations). I then calculated the mean fitness of this population relative to the mean fitness of the original population. Finally, starting at the first update and continuing sequentially, I tested whether the mean relative fitness of the reverted population was less than 1.001; if so, reversion was no longer beneficial at this update. I found that reversion stopped being beneficial sooner at large-effect initial mutations than at small-effect initial mutations (Figure 2.3), indicating that negative epistasis was strongest for large-effect mutations. As I noted before, however, the overall amount of negative epistasis is proportional to the number of mutations accumulated. Because populations that start with lower fitness adapt faster than populations with higher fitness, they accumulate more mutations and therefore have greater total negative epistasis. In this case, negative epistasis is explained by the speed at which compensation proceeds, not by an intrinsic difference in negative epistasis among treatments (i.e., initial fitness of mutant). To determine whether the amount of neg43 800 A Generation at which reversion stopped being beneficial 600 400 200 0 0.1 800 0.3 0.5 0.7 0.9 0.3 0.5 0.7 0.9 B 600 400 200 0 0.1 Initial fitness of mutant (relative to ancestor) Figure 2.3: Generation at which reversion stopped being beneficial at various initial mutant fitness effects. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are the 95% confidence interval of the mean. 44 ative epistasis is different among treatments by accounting for the different speeds at which compensation occurs, I determined the number of mutations accumulated per treatment at which reversion stopped being beneficial. By examining the number of mutations accumulated rather than the number of generations that have elapsed, I controlled for the varying number of mutations that accumulated through time for each treatment. I still found that reversion stopped being beneficial sooner for large-effect initial mutations than for small-effect initial mutations (Figure 2.4), indicating that negative epistasis was strongest for large-effect mutations. To further show that negative epistasis contributed to shortening the window of opportunity for reversion, I estimated the time at which a reversion must appear in order to eventually reach fixation with and without negative epistasis (see Methods). These estimates were calculated using Markov chain simulations and the known probability of reversion and selective coefficient through time. I then compared these results with runs in which reversion did occur. If the estimates calculated in the presence of negative epistasis match the actual runs better than the estimates calculated without epistasis, then I can be sure that negative epistasis was an important contributor to slowing down the rate of reversion. Indeed, I found that the actual runs matched the estimates that were calculated in the presence of negative epistasis (Figure 2.5). Without negative epistasis, a reversion may appear late in evolution and still reach fixation because its selective benefit lasts longer. With epistasis, however, the window of opportunity was small, so a reversion must appear early on if it will fix, which was exactly what I observed in the actual runs that reverted. I have learned that the probability of compensation and reversion depends on the probability of finding these different kinds of mutations. I should therefore expect that the population size, which increases the total number of mutants available, should play an im45 12 P < 0.01, R 2 = 0.14 P < 0.01, R 2 = 0.16 Substitution at which reversion stopped being beneficial 10 8 6 4 2 0 A 0.1 12 0.3 0.5 0.7 0.9 P < 0.01, R 2 = 0.07 P < 0.01, R 2 = 0.02 10 8 6 4 2 0 B 0.1 0.3 0.5 0.7 0.9 Initial fitness of mutant (relative to ancestor) Figure 2.4: Substitution at which reversion stopped being beneficial. at various initial mutant fitness effects. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Lines are linear regressions for each mutant replicate. Error bars are the 95% confidence interval of the mean. 46 Generation at which reversion must appear in order to fix A 300 250 B Expected w/o epistasis Expected w/ epistasis Observed Generation at which reversion must appear in order to fix 250 200 150 150 100 100 50 50 0 Expected w/o epistasis Expected w/ epistasis Observed 200 0 0.1 C 300 0.3 0.5 0.7 0.9 0.1 0.5 0.7 0.9 0.1 D 300 0.3 0.3 0.5 0.7 0.9 300 250 250 200 200 150 150 100 100 50 50 0 0 0.1 0.3 0.5 0.7 0.9 Initial fitness of mutant (relative to ancestor) Figure 2.5: Generation at which reversion must appear in order to fix. (A) Ancestor 1, Mutant 1, (B) Ancestor 1, Mutant 2, (C) Ancestor 2, Mutant 1, and (D) Ancestor 2, Mutant 2. Open circles ( ) represent the first mutant replicate while solid triangles ( ) represent the second mutant replicate. Error bars are the 95% confidence interval of the mean. 47 portant role in the probability of compensation. I know that reversion is more likely to occur if it appears early because compensatory mutations have not had an opportunity to decrease the fitness effect of reversion and because they have not incurred negative epistasis with reversion. In large populations, the opportunity for a reversion to appear early is higher, thus larger populations should revert more often. I next examine the effect of population size on the probability of compensation and reversion. 2.2.3 Effect of population size I evolved the four mutants with relative fitness of 0.5 under four population sizes each: 10, 100, 1,000, and 10,000. Each of these 16 experiments was replicated 100 times, starting with a full population of genetically-identical individuals, and evolved for 10,000 updates (∼ 850 generations). The mutation rate was set to 0.1 mutations per generation for each experiment. I found that the probability of compensatory adaptation was highest at population sizes of 100 and 1,000 but lowest at the extremes of 10 and 10,000 (Figures 2.6A and 2.6B). The probability of reversion was highest at population size of 10,000 (Figures 2.6C and 2.6D). The final fitness of compensated populations was higher the higher the population size (Figures 2.6E and 2.6F). Thus, although intermediate population sizes had the highest probability of compensating, they did not have the highest final fitness; populations with size of 10,000 had the highest fitness. The reason that the probability of compensatory adaptation at population sizes of 10,000 was lower than that at 100 or 1,000 was that the probability of reversion at population sizes of 10,000 was higher than that at 100 or 1,000. However, this did not explain the reason that the probability of compensation at population size of 10 was lower than that at 100 or 1,000 because reversion at population size of 10 was very unlikely. I observed that at population 48 Percent of populations that compensated Ancestor 1 100 Percent of populations that reverted 100 A 80 60 60 40 40 20 20 0 B 80 0 10 100 100 1,000 10,000 10 100 C 80 60 100 1,000 10,000 100 1,000 10,000 D 20 0 10,000 40 20 1,000 60 40 100 80 0 10 Final relative fitness of compensated populations Ancestor 2 1.0 100 1,000 10,000 10 1.0 E 0.9 0.9 0.8 0.8 0.7 0.7 0.6 0.6 0.5 F 0.5 10 100 1,000 10,000 10 Population size Figure 2.6: Effect of population size on compensation. (A) The proportion of runs that underwent compensatory evolution for ancestor 1 and (B) ancestor 2. (C) The proportion of runs that reverted for ancestor 1 and (D) ancestor 2. (E) The proportional increase in fitness for compensatory runs for ancestor 1 and (F) ancestor 2. Open circles ( ) represent the first mutant replicate while solid triangles ( ) represent the second mutant replicate. 49 size 10 many populations decreased in fitness or did not change in fitness. The number of populations that decreased in fitness for each mutant was 27, 45, 36, and 27. The number of those whose fitness stayed the same (within 1%) was 13, 24, 19, and 9 (listed in the same order as above). In contrast, for population size of 100, only 1 of them decreased in fitness out of all 400 runs, and 8 of them stayed the same. Therefore, many of the populations at size 10 that did not compensate were accumulating deleterious mutations due to the small population size. 2.2.4 Effect of mutation rate I evolved the four mutants with relative fitness of 0.5 under five mutation rates each: 0.0001, 0.001, 0.01, 0.1, and 1.0 (mutations per genome per generation). Each of these 16 experiments was replicated 100 times, starting with 1,000 genetically-identical individuals, and evolved for 10,000 updates (∼ 850 generations). The population size was kept at 1,000 throughout each experiment. I found that when the mutation rate was > 0.0001, the probability of compensatory adaptation was high across the mutation rates I tested (Figures 2.7A and 2.7B), except for the second mutant based on the first ancestor (Figure 2.7A, triangle at mutation rate 0.001), in which 69% of the time the population’s fitness did not increase. When the mutation rate was 0.0001, the probability of compensation was low (∼ 20% or less). The probability of reversion increased slightly the higher the mutation rate for the first ancestor (Figure 2.7C), but it was generally low for the second ancestor (Figure 2.7D). The final mean fitness of populations that compensated was generally higher the higher the mutation rate, but for the second ancestor the final fitness was lower at a 1.0 mutation rate than that at 0.1 (Figures 2.7E and 2.7F). 50 Percent of populations that compensated Ancestor 1 100 Percent of populations that reverted 100 A 80 60 60 40 40 20 20 0 B 80 0 0.0001 0.001 100 0.01 0.1 1.0 0.0001 0.001 100 C 80 60 0.01 0.1 1.0 0.01 0.1 1.0 D 20 0 1.0 40 20 0.1 60 40 0.01 80 0 0.0001 0.001 Final relative fitness of compensated populations Ancestor 2 1.0 0.01 0.1 1.0 0.0001 0.001 1.0 E 0.9 0.9 0.8 0.8 0.7 0.7 0.6 0.6 0.5 F 0.5 0.0001 0.001 0.01 0.1 1.0 0.0001 0.001 Mutation rate Figure 2.7: Effect of mutation rate on compensation, reversion, and final population fitness. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are bootstrap 95% confidence intervals of the mean. 51 In the analysis of population size, reversion was highest at the highest population size (10,000) because the total number of mutations that arose was highest, maximizing the likelihood that a revertant mutation appeared. At the highest mutation rate (1.0), the total number of mutations that arose was the same as those in which the population size was 10,000 (10,000 individuals × 0.1 = 1,000 = 1,000 individuals × 1.0). However, I found that reversion was less likely at a mutation rate of 1.0 (compare Figure 2.6C at population size 10,000 with Figure 2.7E at mutation rate 1.0). The reason may be that at the higher mutation rate, in which all types of mutations have a greater chance of arising, the revertant mutation often arises on organisms with other mutations, therefore introducing the possibility of negative epistasis. To test this, I performed a similar analysis as in the initial mutation fitness, where I calculated the number of generations in which a reversion stopped being beneficial in populations that did not eventually revert. The expectation is that the sooner reversions stop being beneficial the stronger that negative epistasis is with other mutations. I found that reversion stopped being beneficial sooner at higher mutation rate than at lower mutation rate (Figure 2.8), indicating that negative epistasis was strongest at higher mutation rate. 2.3 Discussion Our results corroborate previous findings that compensatory adaptation is a common alternative to reversion (Burch and Chao, 1999; Moore et al., 2000; Levin et al., 2000; MaisnierPatin et al., 2002; Estes et al., 2011). Levin et al. (2000) stated that “compensatory evolution establishes an adaptive valley that is difficult to traverse and thus return to the ancestral genotype . . . ” and identified two reasons for this: (1) there are more compensatory mutations than a revertant mutation and (2) in serial transfers, population bottlenecks hinder 52 800 A Generation at which reversion stops being beneficial 600 400 200 0 0.001 0.1 1 0.001 800 0.01 0.01 0.1 1 B 600 400 200 0 Mutation rate Figure 2.8: Generation at which reversion stopped being beneficial at various mutation rates. (A) Ancestor 1 and (B) Ancestor 2. Open circles ( ) and dashed lines represent the first mutant replicate while solid triangles ( ) and solid lines represent the second mutant replicate. Error bars are bootstrap 95% confidence intervals of the mean. 53 the spread of reversions. They found in their experiments that both processes were going on: “the rate of compensatory mutation exceeds that of reversion by at least a factor of 10” and 2/8 experiments reverted using a higher bottleneck but 0/12 experiments reverted using a smaller bottleneck. In my populations, there is no serial passage (like a chemostat), so I do not have the problem of bottlenecks, although populations are limited in size. Yet I see a lot of compensation, meaning that bottleneck reason is not as important as other factors. Instead, I found that negative epistasis between compensatory mutations and a revertant (i.e., absence of the deleterious mutation) prevented the revertant from spreading in a population. 2.3.1 Availability of compensatory mutations Moore et al. (2000) and Sanju´n et al. (2005) found that low-fitness mutants compensated a faster than high-fitness mutants. The reason was that, as Moore et al. (2000) explained, a compensatory mutation in a low-fitness mutant has a higher selective coefficient than in a high-fitness mutant (assuming the compensatory mutation increases fitness relative to the mutant equally for low-fitness and high-fitness mutants). An alternative explanation, which Moore et al. (2000) also discussed but lacked the evidence to support it, was that there may be more compensatory mutations available in low-fitness mutants than in high-fitness mutants. In fact, this was exactly what I found in the digital mutants (Figure 2.1). Our results support Whitlock et al. (2003), who said “when something is broken it is easier to improve than when in is fully functional” and Poon and Chao (2005), who found a positive relationship between the negative effect of a deleterious mutation and the probability of compensation. The reason for this that Poon and Chao (2005) concluded was that “deleterious mutations with large effects on fitness may tend to affect a broader range of phenotypic 54 components. Severely deleterious mutations would therefore generate a larger mutational target for compensatory interactions.” This must be going on in my populations because the more severe the deleterious mutation, the more tasks that are being knocked out. I did not find, as Sanju´n et al. (2005) found, that low-fitness mutants improved in fitness more a than high-fit mutants (Figures 2.2E and 2.2F). 2.3.2 Effect of initial mutant fitness Compensatory mutations depend upon the original deleterious mutation they compensate, such that they may be deleterious when the deleterious mutation is removed, such as when reversion occurs (Schrag et al., 1997; Poon et al., 2005; Levin et al., 2000). Negative epistasis between compensatory mutations and revertant mutations prevented reversion from being beneficial earlier for low-fitness mutants than for high-fitness mutants (Figure 2.3). Reversion in low-fitness mutants may have stopped being beneficial earlier than high-fitness mutants for two reasons. The first reason may be that because in low-fitness mutants there are many more compensatory mutations available than in high-fitness mutants (Figure 2.1), more mutations fix in low-fitness mutants (Moore et al., 2000; Sanju´n et al., 2005), so that a when a reversion appears, it does so in the genetic background with many mutations, in which there are more chances for negative epistasis. The second reason may be due to the fact that large-effect compensatory mutations, which are more likely to fix in low-fitness mutants cause large-effect negative epistasis. I clearly see the latter reason going on: largeeffect compensatory mutations have stronger negative epistasis with the revertant mutation. I cannot, however, exclude the former, and it is likely going on as well. The importance of all this is that as compensatory adaptation proceeds, it becomes increasingly harder for reversion to occur, and populations are obligated to diverge in unique evolutionary paths 55 because typically there are multiple ways to compensate (Paulander et al., 2007). This decreasing probability of reversion as compensatory adaptation proceeds has been inferred in mammalian evolution (Soylemez and Kondrashov, 2012). 2.3.3 Population size Previous studies have generally found that reversion is more common in large populations (Burch and Chao, 1999; Arguello-Astorga et al., 2007). Burch and Chao (1999) found that in two of their large populations (1,000 and 10,000) a single step recovered fitness substantially. The single step for population 1,000 recovered fitness completely and it may have been a revertant mutation, although this is unknown. Smaller populations recovered fitness stepwise, and therefore could not have been revertant mutations, and their smallest populations increased in fitness very slowly. Our results corroborate their findings and those of Sanju´n a et al. (2005) in that the probability of reversion increased with population size and the fitness of populations was greater the greater the population size. In contrast, Maisnier-Patin et al. (2002) observed that compensation happened more readily at greater population size (even in the millions), and reversion was hardly observed. The difference may have been due to differences in mutation rate: viruses and my digital organisms have high mutation rates that may have caused revertants to appear more frequently. Both Burch and Chao (1999) and Maisnier-Patin et al. (2002) found that fitness was highest when the population size was highest, as I did (Figures 2.6E and 2.6F). However, Estes and Lynch (2003) argued that no reversion occurred in their study at their population size of about 10,000, possibly because their mutation rate was very low (4.4 × 10−8 per nucleotide per generation). As Poon and Otto (2000) found, compensatory mutations sometimes helped “freeze” the mutational meltdown that could have occurred in small populations. But there is a 56 limit in population size in which even compensatory adaptation cannot save because of the overwhelming effects of drift (Poon et al., 2005). I found that at population size of 10, many populations decreased in fitness as they accumulated deleterious mutations. Even severely deleterious populations can be compensated, and do so quickly, but little can be done about small populations. 2.3.4 Mutation rate Maisnier-Patin et al. (2002) found that the rate of compensatory adaptation in Salmonella typhimurium was higher in a mutator strain (i.e., higher mutation rate). Perron et al. (2010) arrived at a similar result using a mutator strain of Pseudomonas aeruginosa. Our results corroborate this trend, except at very high mutation rates, where the probability of reversion increased and thus compensation decreased (Figure 2.7). However, I found that the rate of reversion was not as high as expected (based on my treatments with population size) because the window of opportunity for reversion to be be beneficial was very small at high mutation rates (Figure 2.8). The reason was that at high mutation rates, the revertant mutation is likely appear on a genetic background with other mutations that could cause negative epistasis. 2.3.5 Conclusions The fixation of compensatory mutations, which alleviate the negative effects of fixed deleterious mutations, is a common process in which populations recover from deleterious mutations. Reversion is an alternative process, but the competing process of compensation dictates the window of opportunity for which reversion could happen. Populations with large-effect dele- 57 terious mutations have the most number and largest of compensatory mutations available. In addition, large-effect compensatory mutations have strong negative epistasis with the revertant mutation, causing a smaller window of opportunity for reversion the lower the fitness of the population. Larger populations increase the probability of reversion because there are greater chances for reversion to appear within its widow of opportunity. However, although a greater mutation rate has a similar effect, it also shrinks reversion’s window of opportunity because other mutations are also likely to be present, thereby causing negative epistasis. Small populations and low mutation rate slow down adaptation in general, so both compensation and reversion are less likely to occur, but very small populations are likely to decline in fitness. 2.4 2.4.1 Materials and Methods Evolution of ancestors Starting with a digital organism with a genome length of 200, I derived 20 asexual ancestral populations. This initial organism could reproduce but could not perform any tasks. I set the grid (or ‘world’) size to 10,000 individuals and the point mutation rate to 0.1 per genome per generation. Populations were evolved in the default nine-task environment, re-configured to add the bonus for each task performed, rather than multiply its power of two. I allowed 20 replicate asexual populations to evolve for 500,000 updates (∼ 40,000 generations). I then chose two asexual populations whose consensus sequence could perform all nine tasks. The consensus sequences of these two populations served as the ancestral genotypes from which mutants were derived. 58 2.4.2 Construction and evolution of mutants From each ancestor, I generated every possible single mutant (5,000) and chose five pairs of mutants, each pair with the following relative fitnesses: 0.1, 0.3, 0.5, 0.7, and 0.9 (± 0.025). I ensured that the only allele at the mutant locus that could fully recover fitness was the ancestral, revertant, allele. If I could not find mutants with the above conditions for any ancestor, I chose another ancestral genotype that could perform all nine tasks and repeated the method for generating mutants. In total, I obtained 20 mutants: 10 for each of the two asexual ancestors. 2.4.3 Detection of compensatory adaptation To determine whether a population compensated or recovered, I looked at the consensus sequences at the end of 10,000 updates (∼ 850 generations). If the fitness of the consensus sequence reached 99% of the ancestral sequence and the allele at the mutant position changed (to either the ancestral or to something else), then the population was said to have reverted. The reason I also considered non-ancestral alleles as reversions is that substitutions at the mutant locus could represent neutral alleles in the original ancestor; this “effective reversion” is common in some viruses (Arguello-Astorga et al., 2007). If the fitness of the consensus sequence was higher than the mutant sequence and the allele was not the ancestral allele, then the population was said to have compensated (so both full and partial compensation were clumped together). I included substitutions as compensations when the recovery was partial for similar reasons as above: the mutant allele likely has neutral mutations that are effectively equivalent. 59 2.4.4 Estimation of the expected time for reversion I estimated the expected time (in updates) at which a new reversion destined to fixation appeared. To do this, I went through each update t = 100, 200, . . . , 10,000 (my data’s resolution) and stopped at update t with probability p(t), the probability that a new reversion appears between updates t − 100 and t and is destined to fixation. I calculated p(t) as A100 f (s(t))(1 − p(t − 100)), where A100 is the probability that a reversion appears within 100 updates, and f (s(t)) is the probability that a new reversion with selection coefficient s(t) is destined to fixation (methods explained below). The 1 − p(t − 100) is the probability that a reversion destined to fixation did not appear in the previous update. This process was repeated 100 times for each replicate run in the treatment testing the effect of the initial fitness (except for replicates that actually reverted because that would interfere with calculating a reversion’s selection coefficient). I estimated the expected time for two different cases: (1) the revertant’s relative fitness is always 1.0 (i.e., without epistasis) and (2) the revertant’s relative fitness depends on the genetic background on which it appears (i.e., with epistasis). For either case, the probability A100 that a reversion appears within 100 updates will be the same, but the probability of fixation f (s(t)) will be different because s(t) depends on whether there is epistasis. To estimate A100 , I ran 100 Avida experiments where the configuration was identical to the default experiment (except that the initial population consisted of organisms that could not perform any tasks). During the experiments, every time a specific mutation appeared (e.g., the position and allele of a reversion for one of my treatments), I recorded the update at which this happened. (Because the mutation rate was the same for all runs in the treatment that tested the initial fitness, the choice of revertant 60 to record does not matter because the probability that any specific mutation appears is the same for all mutations.) I then binned each recorded update into bins of size 100, counted the number in each bin, and divided that number by 100 (because there were 100 replicate runs). I found that the mean of A100 was 0.1915 with a standard deviation of 0.0443. As mentioned before, the probability of fixation f (s(t)) depends on whether there is epistasis because s(t) depends on this. However, given a specific value of s(t), call it s, I can estimate its fixation probability. To do this, I created every possible mutant of the first ancestor and calculated their fitness, which gave us various values of s for a revertant (calculated as 1 − w(m)/w(a), where w(m)/w(a) is the relative fitness of the mutant). I discarded any mutants with either zero fitness or a fitness greater than the ancestor’s, as these would result in an s value less than or equal to 0 (not a beneficial reversion). Then, for each unique value of s, I populated a new Avida world of 1,000 identical mutants and a single revertant (i.e., the ancestor). (Because several different mutants sometimes had the same s, I picked the first such mutant in the above configuration.) I let each population run in replicates of 100 for 10,000 updates (∼ 850 generations) and zero mutation rate. After removing one outlier, which had 1.0 fixation probability, I fit two line segments to the data such that together they minimized the sum of the squared residuals (the first line was anchored at 0.001 fixation probability for s of 0). I found that when s < 0.28, f (s) = 1.813s (with residual standard deviation 0.09584) and when s ≥ 0.28, f (s) = 0.2414s+0.4382 (with residual standard deviation of 0.0643). I estimated the revertant’s selection coefficient s(t) as wR (t) − w(t), where wR (t) is the revertant’s mean relative fitness and w(t) is the population’s mean relative fitness between updates t and t−100. Because I recorded data at updates 0, 100, 200, . . . , 10,000, I estimated wR (t) and w(t) as the mean value between those at t and t − 100. In the case where there 61 is no epistasis, wR (t) was always 1.0. In the case where there is epistasis, wR (t) was the mean relative fitness of the population where each individual in the population was given the revertant mutation. These analyses were conducted for each replicate run in which reversion did not happen (total of 1,643 runs). 2.5 Acknowledgments This material is based in part upon work supported by the National Science Foundation under Cooperative Agreement No. DBI-0939454. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation. 62 LITERATURE CITED 63 LITERATURE CITED Adami, C. (2006). Digital genetics: unraveling the genetic basis of evolution. Nature Rev. Genet., 7:109–118. Arguello-Astorga, G., Ascencio-Ib´nes, J. T., Dallas, M. B., Orozco, B. M., and Hanleya˜ Bowdoin, L. (2007). High-frequency reversion of geminivirus replication protein mutants during infection. J. Virol., 81:11005–11015. Bj¨rkman, J., Nagaev, I., Berg, O. G., Hughes, D., and Andersson, D. I. (2000). Effects o of environment on compensatory mutations to ameliorate costs of antibiotic resistance. Science, 287:1479–1482. Burch, C. L. and Chao, L. (1999). Evolution by small steps and rugged landscapes in the RNA virus φ6. Genetics, 151:921–927. Chun, S. and Fay, J. C. (2011). Evidence for hitchhiking of deleterious mutations within the human genome. PLoS Genet, 7:e1002240. Elena, S. F., D´vila, M., Novella, I. S., Holland, J. J., Domingo, E., and Moya, A. (1998). a Evolution, 52:309–314. Estes, S. and Lynch, M. (2003). Rapid fitness recovery in mutationally degraded lines of Caenorhabditis elegans. Evolution, 57:1022–1030. Estes, S., Phillips, P. C., and Denver, D. R. (2011). Fitness recovery and compensatory evolution in natural mutant lines of C. elegans. Evolution, 65:2335–2344. Hartl, D. L. and Taubes, C. H. (1996). Compensatory nearly neutral mutations: Selection without adaptation. J. Theor. Biol., 182:303–309. Lande, R. (1994). Risk of population extinction from new deleterious mutations. Evolution, 48:1460–1469. Lenski, R. E., Ofria, C., Collier, T. C., and Adami, C. (1999). Genome complexity, robustness, and genetic interactions in digital organisms. Nature, 400:661–664. Levin, B. R., Perrot, V., and Walker, N. (2000). Compensatory mutations, antibiotic resistance and the population genetics of adaptive evolution in bacteria. Genetics, 154:985–997. Lynch, M., Conery, J., and B¨rger, R. (1995). Mutation accumulation and the extinction of u small populations. The American Naturalist, 146:489–518. Maisnier-Patin, S., Berg, O. G., Liljas, L., and Andersson, D. I. (2002). Compensatory adaptation to the deleterious effect of antibiotic resistance in Salmonella typhimurium. Mol. Microbiol., 46:355–366. Moore, F. B. G., Rozen, D. E., and Lenski, R. E. (2000). Pervasive compensatory adaptation in escherichia coli. Proc. R. Soc. Lond. B, 267:515–522. 64 Ofria, C. and Wilke, C. O. (2004). Avida: a software platform for research in computational evolutionary biology. Artif. Life, 10:191–229. Orr, H. A. (2002). The population genetics of adaptation: the adaptation of DNA sequences. Evolution, 56:1317–1330. Paulander, W., Maisnier-Patin, S., and Andersson, D. I. (2007). Multiple mechanisms to ameliorate the fitness burden of mupirocin resistance in Salmonella typhimurium. Mol. Microbiol., 64:1038–1048. Perron, G. G., Hall, A. R., and Buckling, A. (2010). Hypermutability and compensatory adaptation in antibiotic-resistant bacteria. Am. Nat., 176:303–311. Poon, A. and Chao, L. (2005). The rate of compensatory mutation in the DNA bacteriophage φX174. Genetics, 170:989–999. Poon, A., Davis, B. H., and Chao, L. (2005). The coupon collector and the suppressor mutation: estimating the number of compensatory mutations by maximum likelihood. Genetics, 170:1323–1332. Poon, A. and Otto, S. P. (2000). Compensating for our load of mutations: freezing the meltdown of small populations. Evolution, 54:1467–1479. Presgraves, D. C. (2010). Speciation genetics: search for the missing snowball. Curr. Biol., 20:R1073–R1074. Sanju´n, R., Cuevas, J. M., Moya, A., and Elena, S. (2005). Epistasis and the adaptability a of an RNA virus. Genetics, 170:1001–1008. Schrag, S. J., Perrot, V., and Levin, B. R. (1997). Adaptation to the fitness costs of antibiotic resistance in Escherichia coli. Proc. R. Soc. Lond. B, 264:1287–1291. Soylemez, O. and Kondrashov, F. A. (2012). Estimating the rate of irreversibility in protein evolution. Genome Biol. Evol., 4:1213–1222. Whitlock, M. C. (2000). Fixation of new alleles and the extinction of small populations: drift load, beneficial alleles, and sexual selection. Evolution, 54:1855–1861. Whitlock, M. C., Griswold, C. K., and Peters, A. D. (2003). Compensating for the meltdown: the critical effect size of a population with deleterious and compensatory mutations. Ann. Zool. Fennici, 40:169–183. Wilke, C. O. and Adami, C. (2002). The biology of digital organisms. Trends Ecol. Evol., 17:528–532. 65 Chapter 3 Compensatory adaptation causes rapid incipient speciation 3.1 Introduction Biological speciation is the evolution of reproductive isolating barriers that prevent populations from interbreeding (Coyne and Orr, 2004). One potentially important barrier causes ‘intrinsic postzygotic isolation,’ in which hybrids are sterile or inviable from developmental, physiological, or behavioral abnormalities (Coyne and Orr, 2004). Intrinsic postzygotic isolation is believed to commonly evolve from genetic incompatibilities: negative epistatic interactions among population-specific alleles inherited by the hybrids (Presgraves, 2010). Alleles in a population may arise and spread by natural selection, but the selective pressures involved are often unknown (Schluter, 2009). In order to infer the selective pressures that ultimately caused population-specific alleles involved in genetic incompatibilities, studies using methods in genetic mapping and molecular genetics have helped identify ‘speciation genes’ and their functions (Noor and Feder, 2006; Maheshwari and Barbash, 2011). Many of these genes involve the internal environment of the cell, such as cellular housekeeping, genetic regulation, genetic conflict, and coevolution between nuclear and mitochondrial genomes (Noor and Feder, 2006; Wolf et al., 2010). Interestingly, these genes often appear to have 66 been driven by natural selection (Noor and Feder, 2006), suggesting that adaptation to the internal, genetic environment (as opposed to the external, ecological environment) can lead to intrinsic postzygotic isolation (Phadnis and Orr, 2009; Presgraves, 2010). One class of adaptation to the genetic environment is compensatory adaptation, in which secondary mutations compensate for the effects of accumulated deleterious mutations (Hartl and Taubes, 1996; Burch and Chao, 1999; Moore et al., 2000; Levin et al., 2000; MaisnierPatin et al., 2002; Estes and Lynch, 2003; Estes et al., 2011). Deleterious mutations may accumulate in a population through genetic drift (Lande, 1994; Lynch et al., 1995), hitchhiking with beneficial mutations (Chun and Fay, 2011), transient environmental changes (Bj¨rkman et al., 2000), or spread of selfish genetic elements (Presgraves, 2010). Compeno satory adaptation recovers and maintains the original phenotype via stabilizing selection (i.e., selection against extreme phenotypes), and thus leads to genotypic, not phenotypic, changes (Hartl and Taubes, 1996). Populations undergoing independent compensatory adaptation will therefore diverge genetically, accumulating their own unique set of deleterious and compensatory mutations. Hybrids between such compensated populations would acquire a mismatched set of deleterious and compensatory mutations, exposing genetic incompatibilities that reduce hybrid fitness (Hartl and Taubes, 1996; Orr and Turelli, 2001; Kondrashov et al., 2002; Kulathinal et al., 2004; Landry et al., 2007; Schluter and Conte, 2009; Presgraves, 2010), In this way, compensatory adaptation may lead to intrinsic postzygotic isolation. However, whether compensatory adaptation can lead to postzygotic isolation remains to be tested experimentally. Here I perform such an experiment to answer the following questions: (1) does postzygotic isolation evolve from independent compensatory adaptation? (2) what is the strength of genetic incompatibilities formed? and (3) what is the relative contribution of compensatory mutations and deleterious mutations to postzygotic isolation? 67 Answering these questions requires that I identify both deleterious and compensatory alleles, which involves genetic manipulations that test the allelic effects of each type of mutation. For example, compensatory mutations must not be beneficial in the absence of the deleterious mutations they compensate, and thus they must be tested on their own. Such genetic manipulations, however, are difficult even in model systems, where genetic tools have been greatly advanced. Therefore, I conducted my experiments using the artificial life system Avida (Ofria and Wilke, 2004) (see p. 6), which has been used previously to study various questions in evolution (Lenski et al., 1999, 2003; Chow et al., 2004; Misevic et al., 2006; Elena et al., 2007; Elena and Sanju´n, 2008; Misevic et al., 2010). Avida has enabled research on evolving a genetic systems that would have been difficult in natural systems (Adami, 2006), and the similarities between digital and biological organisms in many evolutionary phenomena have been remarkable (Wilke and Adami, 2002; Adami, 2006). Avida enhances the benefits of microbial systems (i.e., short generation times, considerable replication, easy manipulation and storage of genomes), while being a true instance of evolution of a genetic system, where the generality of evolutionary principles can be tested (Lenski et al., 1999; Elena and Sanju´n, a 2008; Misevic et al., 2006). Specific to this study, Avida allowed us to easily insert deleterious mutations of various effect sizes, carry out thousands of hybridizations, and individually identify compensatory mutations. Using Avida, I isolated mutants with deleterious mutations from a well-adapted ancestor, and I allowed those mutants to evolve in replicate for thousands of generations. I then hybridized compensated populations and measured their fitness to test for postzygotic isolation, and I identified individual compensatory mutations to determine their strength and contribution to postzygotic isolation. I found that (1) postzygotic isolation occurred be68 tween compensated populations, (2) the strength of incompatibility among compensatory mutations was greater than among neutral mutations, and (3) compensatory mutations contributed as much as deleterious mutation to postzygotic isolation. Our results suggest that compensatory adaptation may be an important mechanism by which genetic incompatibilities and thus intrinsic postzygotic isolation evolve. Because I used a non-specific genetic system to test this hypothesis, my results can be generalizable to biological organisms and motivate future tests in biological organisms. 3.2 Results Starting with a sexually-reproducing, haploid digital organism that was not adapted to its environment but could replicate, I allowed three independent populations to evolve for about 250,000 generations under the same environmental configuration. I used the most common genotype of each adapted population as an independent ancestor for all subsequent experiments. From the ancestors, I isolated 472 mutants with 1-5 random irreversible mutations whose combined negative fitness effect was either small (∆W = 0.01-0.1) or large (∆W = 0.10.9). As a control, I also isolated 74 neutral mutants (∆W = 0.0) with the same range of mutations per genome as those in the small-effect and large-effect treatments (Table 3.1). I then allowed populations founded by each mutant (including the controls) to evolve for about 6,000 generations. Populations were evolved in identical environmental conditions as their ancestor because I was interested in compensatory adaptation only. I regarded a population as compensated if its most common genotype (1) had a fitness at least equal to that that of its ancestor, (2) did not acquire mutations that were beneficial on their own, and (3) did not acquire mutations that were deleterious when they first appeared. 69 Treatment Neutral Small Large Ancestor 1 2 3 1 2 3 1 2 3 Mutants 25 25 24 71 199 69 44 44 45 Compensated 10 15 6 9 5 6 Table 3.1: Number of mutants and compensated populations per ancestor for the different treatments 3.2.1 Reproductive isolation via compensation is rapid In nature, subpopulations that split off from their ancestral population may come into secondary contact with their ancestor or with another subpopulation. To model these two scenarios, I performed two types of hybridizations: (1) between compensated populations and their ancestor (‘AC’) and (2) between pairs of compensated populations (‘CC’). I also performed these two types of hybridization on the control populations (i.e., those that evolved starting with neutral mutations). Note that these hybridizations were performed after all experimental evolution had completed, i.e., populations evolved independently. I found that the mean fitness of hybrids for both hybridization types was lower than that of hybrids from control populations (in Fig. 3.1, compare ‘Control’ hybrids with ‘Del. + Comp.’ hybrids). Surprisingly, whether populations compensated for small- or large-effect deleterious mutations did not have a significant effect on mean hybrid fitness (in Fig. 3.1, the 95% bootstrap confidence intervals overlap for ‘Small-effect Del. + Comp.’ and ‘Large-effect Del. + Comp.’). These findings show that intrinsic postzygotic isolation developed faster during compensatory adaptation than during neutral evolution, regardless of the fitness effect size of the initial deleterious mutations. 70 Mean relative fitness A 1.2 1.0 0.8 0.6 0.4 0.2 0.0 Control Del. only Mean relative fitness B Comp. only Del.+Comp. Small effect Neutral effect Comp. only Del.+Comp. Del. only Large effect Comp. only Del.+Comp. Del. only Comp. only Del.+Comp. Small effect Large effect 1.2 1.0 0.8 0.6 0.4 0.2 0.0 Control Neutral effect Del. only Parent 1 Hybrids Parent 2 Figure 3.1: Fitness of hybrids after 25,000 updates of parental evolution. Populations compensated for either small-effect or large-effect deleterious mutations. (A) Hybridizations between compensated genotypes and their ancestor. (B) Hybridizations between pairs of compensated genotypes sharing the same ancestor. ‘Del. Alone’ include only the initial deleterious mutations before compensation, ‘Comp. Alone’ include only the compensatory mutations after compensation, and ‘Del. + Comp.’ include both deleterious and compensatory mutations. ‘Control’ include all mutations accumulated neutrally. Error bars are 95% bootstrap confidence intervals. 71 3.2.2 Compensatory adaptation forms strong genetic incompatibilities Compensated populations acquired 2-10 mutations at the end of the runs, while control populations acquired only 0-1 mutations. Compensated populations thus had greater potential for creating a greater number of genetic incompatibilities than control populations. To determine whether compensated populations created stronger genetic incompatibilities than control populations, I accounted for the number of mutations acquired by each type of population. If hybrids between compensated genotypes have stronger genetic incompatibilities than hybrids between control genotypes, then the rate at which hybrid fitness decays with the number of inherited mutations should be greater for hybrids between compensated genotypes than for hybrids between control genotypes. In other words, the slope of the line relating the number of mutations in hybrids with the hybrid fitness should be greater for hybrids between compensated genotypes than for hybrids between control genotypes. Before I performed this test, however, I generated a new set of 250 control genotypes because my experimental control populations acquired only 0-1 mutations after about 6,000 generations. To generate the new control genotypes, I introduced 2-10 neutral mutations into ancestral genotypes, which was within the range of number of mutations in compensated genotypes. I then fit a least squares linear relationship between the mean number of mutations in hybrids and the hybrid fitness for each treatment (i.e., neutral-, small-, and large-effect for both AC and CC hybrids). I found that the slope of this linear relationship was significantly greater for hybrids between compensated genotypes than for hybrids between control genotypes (Fig. 3.2B), but was not significantly greater for AC hybrids 72 (Fig. 3.2A). Our findings suggest that genetic incompatibilities involving deleterious and compensatory mutations in hybrids were stronger than those among neutral mutations. 3.2.3 Both deleterious and compensatory mutations contribute to reproductive isolation Although I have shown that genetic interactions involving deleterious and compensatory mutations may form strong genetic incompatibilities rapidly, I have yet to quantify the relative contributions of each type of mutation to the formation of genetic incompatibilities. At first glance, Fig. 3.1 may suggest that compensatory mutations contributed little to reproductive isolation because hybrids between genotypes before compensation (‘Del. only’) had fitnesses as low as hybrids after compensation (‘Del. + Comp.’). However, hybrids between compensated genotypes in which deleterious mutations were reverted to the ancestral state also had low fitnesses (in Fig. 3.1, compare ‘Comp. only’ hybrids to ‘Del. + Comp.’ hybrids), suggesting that compensatory mutations also contributed to postzygotic isolation. To establish more directly the extent to which incomplete sets of deleterious and their corresponding compensatory mutations contributed to postzygotic isolation, I generated genotypes with different proportions of deleterious and compensatory mutations. I generated such genotypes by creating every possible hybrid between a compensated genotype and its ancestor. For each genotype, I then measured its fitness and calculated the proportion of deleterious and compensatory mutations relative to the compensated parental genotype. I performed this analysis on all compensated genotypes for both the small-effect and largeeffect treatments. I found that, unless genotypes contained either all or none of their parent’s 73 1.2 AC 1.0 0.8 0.6 Slope 0.4 0.0 Mean hybrid fitness 0.2 −0.2 A Nt. Sm. Lg. 0.0 0 1 2 3 4 5 6 7 1.2 8 CC 1.0 0.8 0.6 a 0.0 b b Slope 0.4 0.2 −0.2 B Nt. Sm. Lg. 0.0 0 1 2 3 4 5 6 7 8 Mean number of mutations Neutral Small-effect Large-effect Figure 3.2: Strength of genetic incompatibilities in hybrids. (A) Least squares linear fit of the mean number of mutations in hybrids between ancestor and compensated genotypes (AC) against their mean fitness. (B) Same as (A) except for hybrids between pairs of compensated genotypes (CC). Insets: Means and 95% bootstrap confidence intervals of each slope where shared letters indicate overlapping confidence intervals. 74 deleterious and compensatory mutations, their fitness was low (Fig. 3.3), suggesting that both deleterious and compensatory mutations contributed to postzygotic reproductive isolation. 3.3 Discussion Most genes involved in intrinsic postzygotic isolation, i.e., hybrid sterility or inviability due to developmental or physiological abnormalities, show strong signatures of positive selection (Presgraves, 2010). Surprisingly, many of these genes are not adaptations to the external, ecological environment but to an impaired internal, genetic environment (see Table 3.1 in (Presgraves, 2010)). An example of adaptation to an impaired genetic environment is adaptive compensation of deleterious mutations (Hartl and Taubes, 1996; Burch and Chao, 1999; Moore et al., 2000; Levin et al., 2000; Maisnier-Patin et al., 2002; Estes and Lynch, 2003; Estes et al., 2011). Populations undergoing independent compensatory adaptation will diverge genetically and may form genetic incompatibilities, causing intrinsic postzygotic isolation (Orr and Turelli, 2001; Kondrashov et al., 2002; Kulathinal et al., 2004; Coyne and Orr, 2004; Landry et al., 2007; Schluter and Conte, 2009; Presgraves, 2010). Under this scenario, I asked: how rapidly does postzygotic isolation evolve? what is the strength of genetic incompatibilities? and what are the relative contributions of compensatory and deleterious mutations to isolation? Using an artificial life system, I found that postzygotic isolation due to compensation was rapid: hybrids between compensated populations had significantly lower fitness than hybrids between populations that did not undergo compensation, regardless of the effect size of the initial deleterious mutations. I also found that compensatory adaptation formed stronger genetic incompatibilities: the rate at which hybrid fitness decayed with the number 75 Figure 3.3: Fitness of hybrids with intermediate parental contributions of deleterious and compensatory mutations. (A) Mean fitness of hybrids between the ancestor and compensated parents that inherit the specified percent of small-effect deleterious and compensatory mutations. (B) Same as (A) except for large-effect deleterious mutations. 76 of mutations was significantly greater for hybrids between compensated populations. Finally, I found that both deleterious and compensatory mutations contributed to postzygotic isolation: hybrids with different proportions of compensatory and deleterious mutations were unfit unless all or none of both types of mutations were present. Two important implications to the genetics of postzygotic isolation can be drawn from my findings. First, evidence of genotypic diversification between species may not correlate with phenotypic diversification. This is because compensatory adaptation can build up genetic differences without altering phenotypic characteristics. In fact, compensatory adaptation may act under stabilizing selection to maintain phenotypes despite continual accumulation of deleterious mutations (Hartl and Taubes, 1996). Second, evidence of positive selection at loci that contribute to speciation (‘speciation genes’) may not always be the result of diversification due to ecological adaptation but instead be the footprint of compensatory adaptation. Our results corroborate the view that adaptation to the internal, genetic environment may be important in the development of intrinsic postzygotic isolation (Presgraves, 2010; Phadnis and Orr, 2009). In describing his mechanism of founder speciation, Ernst Mayr stated that “. . . the mere change of the genetic environment may change the selective value of a gene very considerably” (Templeton, 2008). In cases where founder effects cause the fixation or high frequency of deleterious mutations, compensatory adaptation may provide a mechanism by which genes with altered selective values change in allele frequency. Furthermore, Templeton’s model of genetic transilience recognizes that founder populations may be affected by genetic drift while new mutations and recombination increase genetic variation (Templeton, 2008). Genetic drift alone can raise the frequency of deleterious mutations in a founder population while increased genetic variation may introduce new compensatory mutations on which selection 77 can act. Our study using experimental evolution with populations of digital organisms suggests that such compensatory mutations can rapidly generate postzygotic barriers, and therefore supports a novel genetic mechanism for models of founder or peripatric speciation. Our study took advantage of the recent implementation of sexual reproduction in the artificial life platform Avida (Misevic et al., 2006), and extended its application as proof of principle to the field of speciation. I demonstrated that Avida is a useful tool to complement other approaches in speciation research as it allows for the direct observation of evolution and reproductive isolation in action. Furthermore, my conclusions will motivate additional research into compensatory adaptation as a viable mechanism for speciation, to be further explored in biological systems. The notion that organisms construct or choose their own microhabitats (Lewontin, 2000) suggests that organisms are somewhat resilient to changes in the external environment. This reduced emphasis on the external environment in evolution is supported by my conclusion that environmental differences between allopatric populations are not essential for genetic diversification. 3.4 3.4.1 Materials and Methods Strains and experimental conditions The starting digital organism I used was the ‘default’ sexually-reproducing organism in Avida, with a genome length fixed to 80 instructions. The population size was set to 10,000 organisms, and the copy mutation probability was set to 0.0005 per instruction (i.e., 0.04 per genome per generation). Other types of mutations, such as insertions or duplications, were not permitted because they may disrupt homologous recombination. Digital organisms were configured to reproduce sexually, and the two recombinant offspring were set to re78 place random individuals in the population when no free space was available. The ‘diverse’ environment rewarded the nine tasks commonly used in Avida experiments (Lenski et al., 2003). The small genome size of digital organisms caused reversions to be common during compensatory adaptation in preliminary runs. To guarantee that genotypic changes were due to mutations at secondary loci—as has been observed in biological organisms (Burch and Chao, 1999; Estes et al., 2011)—I prevented reversions from occurring. 3.4.2 Isolation of mutants To generate mutants for the small-effect and large-effect treatments, I first generated sets of 10,000 random mutant genotypes for each evolved ancestor until either four mutants with the desired number of mutations and effect size were found, or until 100 million mutants had been searched. For the small-effect treatment, effect sizes of 0.01, 0.02, 0.03, 0.04, 0.05, 0.06, 0.07, 0.08, 0.09, and 0.1 (each with up to a 0.0049 deviation) were identified for genotypes carrying 1, 2, 3, 4, or 5 mutations. For the large-effect treatment, mutants with effect sizes of 0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.8, and 0.9 (with up to a 0.049 deviation) for each mutation number of 1, 2, 3, 4, and 5 were identified. (Effect size is the difference between the fitnesses of the ancestral and the mutated genotype, relative to the ancestral genotype.) I isolated a total of 339 small-effect mutants (after running the above procedure twice) and 133 largeeffect mutants. I implemented a different searching procedure for neutral mutations because the probability of several random mutations resulting in a neutral genotype was very low. To find neutral mutants, I first generated 10,000 single-mutants and selected the first one that was neutral (i.e., relative fitness of exactly 1.0). This procedure was repeated starting with the neutral mutant until the desired number of mutations was reached or until the recursion was exhausted. I isolated a total of 74 neutral mutants. 79 3.4.3 Compensated populations I regarded a population as compensated if its most common genotype (1) reached a fitness of at least 1.0 relative to its ancestor, (2) did not acquire mutations that were beneficial on their own, and (3) did not acquire mutations that were deleterious when they first appeared. To determine condition (1), the most common genotype at the end of a run was isolated and its fitness relative to its ancestor was measured. If its relative fitness was equal to or greater than 1.0, then the population was considered compensated. To determine condition (2), the fitness effect of each secondary mutation in the most common genotype of the population was tested in the genetic background of the ancestor. If any ‘transformant’ had a relative fitness above 1.0, then that population was not considered as compensated because mutations were beneficial on their own (i.e., generally beneficial, not compensatory); otherwise, the population was considered compensated. To determine condition (3), I sequentially examined the most common genotype of the population about every three generations. If the gain of any mutation resulted in a lower fitness than the genotype at the previous third generation, then that population was not considered as compensated; otherwise, the population was considered compensated. 3.4.4 Hybridization method Hybrids were created by the same method in which Avida creates recombinants during sexual reproduction: the genetic region between two crossover points were exchanged between two parents to produce two offspring. Hybridizations, whether between compensated populations and their ancestor or between pairs of compensated populations, involved creating every possible hybrid between the most common genotypes of each population. 80 3.4.5 Statistics To compare the fitness of hybrids among treatments (Fig. 3.1), I determined whether their 95% bootstrap confidence intervals overlapped. For each treatment, bootstrap replicates were set to contain the same number of samples per ancestor—established as the mean number of observed samples per ancestor—in order to minimize any potential bias from the fact that ancestors yielded different numbers of compensated populations. All bootstraps contained 10,000 replicates. To determine the linear relationship between the number of mutations in hybrids and their fitness (Fig. 3.2), I fit a linear least squares model to each bootstrap replicate. I compared the 95% bootstrap confidence intervals of the slopes among treatments to establish the relative effect of mutation number on hybrid fitness. Statistical analyses were performed in R (ver. 2.8.1). 3.5 Acknowledgments I thank R. E. Lenski, D. W. Schemske, and C. Ofria, I. Dworkin, and A. Gerstein for helpful discussion and comments on the manuscripts. This work was supported by the Quantitative Biology Initiative at Michigan State University (MSU) and the BEACON Center for the Study of Evolution in Action. Computational experiments were made possible by the High Performance Computing Center at MSU. This material is based in part upon work supported by the National Science Foundation under Cooperative Agreement No. DBI-0939454. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation. 81 LITERATURE CITED 82 LITERATURE CITED Adami, C. (2006). Digital genetics: unraveling the genetic basis of evolution. Nature Rev. Genet., 7:109–118. Bj¨rkman, J., Nagaev, I., Berg, O. G., Hughes, D., and Andersson, D. I. (2000). Effects o of environment on compensatory mutations to ameliorate costs of antibiotic resistance. Science, 287:1479–1482. Burch, C. L. and Chao, L. (1999). Evolution by small steps and rugged landscapes in the RNA virus φ6. Genetics, 151:921–927. Chow, S. S., Wilke, C. O., Ofria, C., Lenski, R. E., and Adami, C. (2004). Adaptive radiation from resource competition in digital organisms. Science, 305:84–86. Chun, S. and Fay, J. C. (2011). Evidence for hitchhiking of deleterious mutations within the human genome. PLoS Genet, 7:e1002240. Coyne, J. A. and Orr, H. A. (2004). Speciation. Sinauer Associates, Sunderland, MA. Elena, S. F. and Sanju´n, R. (2008). The effect of genetic robustness on evolvability in a digital organisms. BMC Evol. Biol., 8:284. Elena, S. F., Wilke, C. O., Ofria, C., and Lenski, R. E. (2007). Effects of population size and mutation rate on the evolution of mutational robustness. Evolution, 61:666–674. Estes, S. and Lynch, M. (2003). Rapid fitness recovery in mutationally degraded lines of Caenorhabditis elegans. Evolution, 57:1022–1030. Estes, S., Phillips, P. C., and Denver, D. R. (2011). Fitness recovery and compensatory evolution in natural mutant lines of C. elegans. Evolution, 65:2335–2344. Hartl, D. L. and Taubes, C. H. (1996). Compensatory nearly neutral mutations: Selection without adaptation. J. Theor. Biol., 182:303–309. Kondrashov, A. S., Sunyaev, S., and Kondrashov, F. A. (2002). Dobzhansky-muller incompatibilities in protein evolution. Proc. Natl. Acad. Sci. USA, 99:14878–14883. Kulathinal, R. J., Bettencourt, B. R., and Hartl, D. L. (2004). Compensated deleterious mutations in insect genomes. Science, 306:1553–1554. Lande, R. (1994). Risk of population extinction from fixation of new deleterious mutations. Evolution, 48:1460–1469. Landry, C. R., Hartl, D. L., and Ranz, J. M. (2007). Genome clashes in hybrids: insights from gene expression. Heredity, 99:483–493. Lenski, R. E., Ofria, C., Collier, T. C., and Adami, C. (1999). Genome complexity, robustness and genetic interactions in digital organisms. Nature, 400:661–664. 83 Lenski, R. E., Ofria, C., Pennock, R. T., and Adami, C. (2003). The evolutionary origin of complex features. Nature, 423:139–144. Levin, B. R., Perrot, V., and Walker, N. (2000). Compensatory mutations, antibiotic resistance and the population genetics of adaptive evolution in bacteria. Genetics, 154:985–997. Lewontin, R. C. (2000). The triple helix: gene, organism, and environment. Harvard University Press, Cambridge, MA. Lynch, M., Conery, J., and B¨rger, R. (1995). Mutational meltdowns in sexual populations. u Evolution, 49:1067–1080. Maheshwari, S. and Barbash, D. A. (2011). The genetics of hybrid incompatibilities. Annu. Rev. Genet., 45:331–355. Maisnier-Patin, S., Berg, O. G., Liljas, L., and Andersson, D. I. (2002). Compensatory adaptation to the deleterious effect of antibiotic resistance in Salmonella typhimurium. Mol. Microbiol., 46:355–366. Misevic, D., Ofria, C., and Lenski, R. E. (2006). Sexual reproduction reshapes the genetic architecture of digital organisms. Proc. R. Soc. B, 273:457–464. Misevic, D., Ofria, C., and Lenski, R. E. (2010). Experiments with digital organisms on the origin and maintenance of sex in changing environments. J. Hered., 101 Suppl 1:S46–54. Moore, F. B. G., Rozen, D. E., and Lenski, R. E. (2000). Pervasive compensatory adaptation in escherichia coli. Proc. R. Soc. Lond. B, 267:515–522. Noor, M. A. F. and Feder, J. L. (2006). Speciation genetics: evolving approaches. Nat. Rev. Genet., 7:851–861. Ofria, C. and Wilke, C. O. (2004). Avida: A software platform for research in computational evolutionary biology. Artif. Life, 10:191–229. Orr, H. A. and Turelli, M. (2001). The evolution of postzygotic isolation: accumulating dobzhansky-muller incompatibilities. Evolution, 55:1085–1094. Phadnis, N. and Orr, H. A. (2009). A single gene causes both male sterility and segregation distortion in Drosophila hybrids. Science, 323:376–379. Presgraves, D. C. (2010). The molecular evolutionary basis of species formation. Nat. Rev. Genet., 11:175–180. Schluter, D. (2009). Evidence for ecological speciation and its alternative. Science, 323:737– 741. Schluter, D. and Conte, G. L. (2009). Genetics and ecological adaptation. Proc. Natl. Acad. Sci. USA, 106:9955–9962. 84 Templeton, A. R. (2008). The reality and importance of founder speciation in evolution. BioEssays, 30:470–479. Wilke, C. O. and Adami, C. (2002). The biology of digital organisms. Trends Ecol. Evol., 17:528–532. Wolf, J. B. W., J., L., and Backstr¨m (2010). Speciation genetics: current status and o evolving approaches. Proc. R. Soc. B, 365:1717–1733. 85 Chapter 4 Ecological and mutation-order speciation in digital organisms 4.1 Introduction Reproductive isolation between populations often evolves as a byproduct of independent adaptation to new environments (Coyne and Orr, 2004; Schluter, 2009; Sobel et al., 2010). When these environments’ selective pressures are different, divergent selection can cause populations to acquire different, often incompatible, alleles. This divergent process can generate reproductive isolation both in nature (reviewed in Rundle and Nosil, 2005; Schluter, 2009) and in laboratory experiments (Dettman et al., 2007, 2008; reviewed in Rice and Hostert, 1993 and Fry, 2009). Complete reproductive isolation due to this process is known as ‘ecological speciation’ (Schluter, 2009). On the other hand, if the environments’ selective pressures are similar or identical (parallel or uniform selection), populations may diverge genetically by the chance fixation of different alleles. Although laboratory experiments and theoretical studies suggest that such process may lead to ‘mutation-order speciation’ (Schluter, 2009; Nosil and Flaxman, 2011), its effectiveness in generating reproductive isolation compared to ecological speciation is unknown. 86 The main purpose of this study is to directly compare the strength of reproductive isolation generated by ecological and mutation-order processes. Specifically, I measure both the degree of postzygotic isolation (i.e., hybrid inviability) as well as the amount of genetic divergence between populations evolved under either different environments or the same environments. Because there is a higher chance of parallel evolution when environments are similar (Schluter and Conte, 2009), I expect that postzygotic isolation and genetic divergence under a mutation-order process will be weaker than under an ecological process. I also examine the effect of migration on both ecological and mutation-order processes. Migration between populations increases the chance of gene flow, which often slows genetic (and thus adaptive) divergence, although gene flow can also promote divergence (Garant et al., 2006; R¨s¨nen and Hendry, 2008). I vary the amount of migration between populations aa under both ecological and mutation-order processes, from allopatry to sympatry. I expect that migration will have a stronger negative effect on the evolution of reproductive isolation under mutation-order scenarios because, under uniform selection, an adaptive mutation that arises in one population is also selectively favored in the other (Schluter, 2009; Nosil and Harmon, 2009; Nosil and Flaxman, 2011). I also examine how the environments’ dimensionality (i.e., number of selective pressures) affects the strength of reproductive isolation for both ecological and mutation-order processes. In high-dimensional environments, there are more opportunities for populations to adapt in different ways (Rice and Hostert, 1993; Nosil et al., 2009), which may lead to stronger reproductive isolation for both ecological and mutation-order processes. High dimensionality, however, may also constrain speciation when trade-offs among adaptive traits hinder adaptation and therefore decrease the probability of reproductive isolation. 87 Finally, I examine one possible cause for differences in hybrid fitness between ecological and mutation-order processes. As populations adapt to their local environments independently, there is no guarantee that alleles acquired in one population will interact positively, or even neutrally, in the hybrid with alleles acquired in the other population (Coyne and Orr, 2004). Negative interactions between alleles from two populations are known as Dobzhansky-Muller incompatibilities (DMIs), and both ecological and mutation-order process may cause them to form and thus lead to postzygotic isolation (Schluter, 2009). It is unknown, however, whether ecological or mutation-order speciation differ in their propensity to produce DMIs. Although a mutation-order process may be expected to have fewer DMIs initially than an ecological process because uniform selection selects for the same alleles, such expectation diminishes the more populations diverge (Schluter, 2009). In this study, I use the software Avida (see p. 6 to carry out my experiments. Avida (Ofria and Wilke, 2004) is an artificial life research platform where digital organisms evolve due to genetic variation, inheritance, and differential reproduction (see Methods). Avida has been used previously in a wide range of ecological and evolutionary studies (e.g., Lenski et al., 2003; Chow et al., 2004; Elena et al., 2007; Ostrowski et al., 1997; Misevic et al., 2010). There are several reasons for using digital organisms to study evolution: I can observe millions of generations in a few days, conduct hundreds of replicate experiments, easily manipulate genomes, and accurately measure fitness. Digital organisms in Avida are not meant to specifically mimic the details of real biological organisms. Instead, digital organisms have a unique genetic system (see Methods). Despite these differences, the general principles that make evolution possible are still the same, which allows Avida to be used to test the generality of evolutionary theories and hypotheses. Indeed, several evolutionary properties have been found to be remarkably similar to that of biological organisms (Wilke and Adami, 88 2002; Adami, 2006) (e.g., the distribution of mutational effects, the types of epistasis, and the genetic architecture of sexual organisms). Digital organisms improve on simple twolocus models of speciation because in Avida, traits are complex, involving multiple loci and epistatic interactions among alleles (Lenski et al., 1999). 4.2 4.2.1 Methods Experimental design The Avida configuration files used to run my experiments are available in the Dryad Data Repository. To generate the ancestral population, I founded a population with an organism that could replicate but could not perform any tasks. I then let this population evolve under the default nine-task environment for 500,000 updates (∼ 42,000 generations). An ‘update’ is a measurement of time in Avida, increasing by one each time organisms execute 60 instructions (on average). For the evolution of the ancestral population, I set the maximum population size to 10,000 individuals. The length of the genome was set to 200 instructions, and to ensure homologous recombination during sexual reproduction, the genome length was fixed. The mutation rate was set to 0.1 mutations per genome per generation. I then set up four treatments (described below), which I call ‘drift,’ ‘ecological,’ ‘mutation-order 1,’ and ‘mutation-order 2.’ For each treatment, the population size was set to 2,000 individuals and divided into two demes, each of size 1,000. In the drift treatment, both demes’ environments were the same as the ancestral (environment ‘A’). For each remaining treatments, I set up two subtreatments (described below): low dimensionality and high dimensionality. For this study, dimensionality refers to the number of tasks for which that environment rewards organisms for performing such tasks. In the ecological treatment, 89 the demes’ environments were different from each other and different from the ancestral (environments ‘1L’ and ‘2L’ for low dimensionality and environments ‘1H’ and ‘2H’ for high dimensionality). In the mutation-order 1 treatment, the demes’ environments were the same as each other but different from the ancestral (environment ‘1L’ for low dimensionality and environment ‘1H’ for high dimensionality). Similarly, in the mutation-order 2 treatment, the demes’ environments were the same as each other but different from the ancestral (environment ‘2L’ for low dimensionality and environment ‘2H’ for high dimensionality). The specific tasks that were rewarded in each environment remained the same for the rest of this study. The number of tasks for the low and high environments were chosen as two extremes: two tasks for low dimensionality and the maximum of 34 (i.e., 68 possible tasks divided randomly into two demes) for high dimensionality. The specific tasks for environments 1L and 2L were chosen at random from tasks that were known to evolve within 10,000 updates in preliminary runs. Environments 1L and 2L shared no tasks; similarly, environments 1H and 2H shared no tasks. The specific tasks rewarded in each environment are part of the Avida configuration files, which are available in the Dryad Data Repository. Each treatment was replicated 20 times with a different random sample of 2,000 organisms (1,000 per deme) from the ancestral population. Successive random samples were reused for each treatment, so that the genotypes in replicate n of a treatment were the same as the genotypes in replicate n of another treatment. I ran each replicate for 10,000 updates (∼ 850 generations). For each run, the entire population of organisms was saved every 100 updates (∼ 8.5 generations). To examine the effect of gene flow, each replicate was run under eight migration rates (for the entire length of the run): 0.0 (allopatry), 0.00001, 0.0001, 0.001, 0.01, 0.05, 0.1, and 0.5 (sympatry). Migration rate is the probability of an offspring being born in a deme different than its parents; the migrating offspring was placed in a random 90 location in the other deme (i.e., there were no hybrid zones) but the parents remained in their own demes. In all, there were 1,120 runs. I measured the overall strength of selection in each environment (1L, 2L, 1H, and 2H) at the end of each ‘ecological’ replicate run. First, for every organism in a population that could perform at least one task, I counted the number of tasks it could perform and calculated its fitness relative to the mean. I then used linear regression on the relationship between the number of tasks an organism could perform and its relative fitness. The slope of this line is the strength of selection (Conner and Hartl, 2004). I report these results here, as they are part of the environment in which populations evolved. For environments 1L and 2L, the mean strengths of selection were 0.3440 (0.3375–0.3562, 95% bootstrap C.I.) and 0.3353 (0.3341–0.3366), respectively. For environments 1H and 2H, the mean strengths of selection were 0.2045 (0.1852–0.2248) and 0.1899 (0.1762–0.2041), respectively. I discuss the strength of selection in the Discussion. 4.2.2 Postzygotic reproductive isolation In this study, I focus on the evolution of postzygotic reproductive isolation. To measure the strength of postzygotic isolation for each treatment, I first selected 1,000 random pairs of organisms (one from each deme) and created one hybrid per pair at the end of each replicate run. I then calculated the fitness of each hybrid as the mean fitness relative to each parent. Finally, I compared the mean hybrid fitnesses for each treatment—the lower this fitness, the stronger the isolation. Note that hybrids were created after the experiments were finished; no hybrids were put back into the population. Two types of hybridizations were performed for creating hybrids after the populations had evolved. The first followed the method used in Avida for sexual reproduction (and 91 the way in which all my experimental populations experienced): a randomly-sized genomic region starting at a random locus was chosen (both random numbers came from a uniform distribution), and two recombinant offspring were created by exchanging the genetic region of one parent with the other (two-point crossover). I randomly chose one of the two offspring as the hybrid. I also performed a more fine-scaled hybridization method, where each locus of a hybrid had the same probability (0.5) of it coming from either one or the other parent. This method effectively increased the number of crossover points up to 200 and the number of regions that can be exchanged up to 100. I used this multiple-point crossover method to break apart coadapted gene complexes, following the same logic that researchers use when carrying out parental backcrosses or intercrosses between hybrids (e.g., Li et al., 1997; Burton et al., 1999). Multiple crossover points can expose incompatible gene complexes between species, revealing patterns of divergence that would be difficult to detect with recombination at only two crossover points. 4.2.3 Genetic divergence To quantify the homogenizing effects of gene flow, I calculated the genetic divergence between each replicate pair of demes under 0.0 and 0.01 migration for each treatment. Genetic divergence was measured as the fixation index FST = 1 − HS /HT , where HS is the mean heterozygosity of each deme and HT is the heterozygosity of both demes treated as one population (Hartl and Clark, 1997, p. 118). The heterozygosity of a deme is the mean heterozygosity at all loci. The heterozygosity at a locus is H = 1 − n 2 i=1 xi , where n is the number of alleles segregating at that locus and xi is the frequency of the ith allele (Gillespie, 2004, p. 15). FST values between 0 and 0.05 would indicate little or no genetic divergence between two demes (Hartl and Clark, 1997, p. 118). I expect that under zero migration FST 92 values will be significantly higher than those under the 0.01 migration rate. Significance among treatments was determined by comparing their 95% confidence intervals of the mean FST . Each confidence interval was estimated by calculating 10,000 means of random samples (with replacement) of the FST values from the 20 replicates (i.e., each sample contained 20 FST values). The interval between 2.5% and 97.5% of means defined the confidence interval. To test whether gene flow causes the same mutations—specifically those involved in performing a task—to fix under mutation-order speciation, I carried out a two-step process to identify and map such mutations. First, to determine whether the fixed mutations in a deme were necessary to perform a task, I reverted each locus, one by one, of the deme’s consensus sequence to the ancestral state. If any reversion eliminated the ability to perform a task, then the allele at that locus must be important for that task. I ignored loci in which a reversion caused complete inviability of the organism, as these loci were involved in more than just task performance. Second, I aligned the consensus sequences of each pair of demes under 0.0 and 0.01 migration and highlighted the mutations I found above. 4.2.4 Hybrid phenotypes To examine a possible cause for differences in hybrid fitness between ecological and mutationorder processes, I counted the number of times that hybrids had low fitness due to Dobzhansky-Muller incompatibilities (DMIs). Under two-point crossover recombination, a hybrid is made up of two parental components, which I call C1 and C2 . If C1 or C2 contains the instructions to perform a task but the full hybrid cannot perform that task, C1 and C2 must interfere with one another through at least one DMI. To determine whether a hybrid had low fitness due to DMIs, I constructed two genotypes by making two copies of the hybrid, where I replaced C2 in the first copy, and C1 in the second copy, with the corresponding ancestral 93 genetic region. In this way, I constructed two ‘component’ genotypes, where each parental component was isolated in the genetic background of the ancestor. I then determined the tasks that these component genotypes as well as the original hybrid could perform. If either component genotype could perform a task but the hybrid could not, then at least one DMI was present (Figure 4.1). I performed this analysis on 1,000 hybrids per replicate in both the ecological and mutation-order treatments under zero migration, low dimensionality, and two-point crossover recombination. Because in this system an organism’s fitness is largely determined by the number and type of tasks it can perform (i.e., its phenotype), I identified the tasks that could be performed by each hybrid for both ecological and mutation-order processes. Note that this analysis is independent of the environment because an organism may have the ability to perform a task even if the environment does not reward for it. For simplicity, I focused only on the zero migration, low dimensionality set of treatments that were hybridized with a two-point crossover. For the ecological treatment, hybrids were categorized by the number tasks they could perform: (‘0-0’) no tasks in either environment, (‘1-0’) one task in one environment but none in the other, (‘1-1’) one task in each environment, (‘2-0’) two tasks in one environment but none in the other, (‘2-1’) two tasks in one environment and one in the other, and (‘2-2’) two tasks in both environments. For the mutation-order treatment, hybrids were categorized by the tasks they could perform: (‘None’) no tasks, (‘1’) task 1, (‘2’) task 2, and (‘1 and 2’) both tasks. For those hybrids that could perform both tasks in the mutation-order treatment, I determined the tasks that each hybrid’s parental components could perform. I categorized these parental components as (‘0,0’) no parental component performs any task, (‘1,0’) one parental component performs one task but the other none, (‘1,1’) each parental component performs a different task, and (‘2,*’) at least one parental component performs 94 Hybrid cannot perform task BUT OR Component genotype 1 can perform task Component genotype 2 can perform task IMPLIES Incompatibility between C1 and C2 C1 C2 Ancestor Figure 4.1: Method to determine whether a hybrid contains genetic incompatibilities. A hybrid is composed of two parental components, C1 and C2 . Note that a parental component is only the parental region inherited by the hybrid; it is not the complete parent. If the hybrid cannot perform the task but either parental component can, then there must be at least one incompatibility between the components. 95 both tasks. This analysis will reveal the reason, at the phenotypic level, for differences in hybrid performance between ecological and mutation-order processes. Four replicates from the ecological treatment, one replicate from the mutation-order 1 treatment, and six replicates from the mutation-order 2 treatment were removed from the analysis above. In the ecological treatment, the removed replicates contained parents that could fortuitously perform a task of the other environment (even though there was no selective pressure for that task), and thus it would be unclear from which parent the task was inherited by the hybrids. In the mutation-order treatments, the removed replicates contained parents that could not perform both tasks and, therefore, was the reason that some of the hybrids were unfit. 4.3 4.3.1 Results Postzygotic reproductive isolation When hybrids between the evolved demes were created by recombining a single genetic region (‘two-point crossover’), reproductive isolation between demes that adapted to different environments (ecological treatment) was considerably stronger than reproductive isolation between demes that adapted to the same environment (mutation-order treatment) (Figs. 4.2A and 4.2B). With zero migration, for instance, reproductive isolation in the ecological treatment was more than twice as strong than in the mutation-order treatment. There was no reproductive isolation between demes evolving neutrally in the ancestral environment (drift treatment): the mean hybrid fitness was > 0.99 at all migration rates. 96 1.0 0.8 0.8 0.6 0.6 0.4 0.4 0.2 Mean hybrid fitness 1.0 0.2 0.0 A 0.0 Low dimensionality, two-point crossover 0.00001 0.0001 0.001 0.01 0.05 0.1 0.0 0.5 B 0.0 1.0 0.1 0.5 0.4 0.2 0.05 0.6 0.4 0.01 0.8 0.6 0.00001 0.0001 0.001 1.0 0.8 High dimensionality, two-point crossover 0.2 0.0 C 0.0 Low dimensionality, multiple-point crossover 0.00001 0.0001 0.001 0.01 0.05 0.1 0.0 0.5 D 0.0 High dimensionality, multiple-point crossover 0.00001 0.0001 0.001 0.01 0.05 0.1 0.5 Migration rate Genetic drift Ecological Mutation−order 1 Mutation−order 2 Figure 4.2: Mean fitness of hybrids (y-axis) between populations evolved under various migration rates (x-axis) and different treatments (markers and lines): the ancestral environment (‘Genetic drift’), different new environments (‘Ecological’), or the same new environment (‘Mutation-order 1’ and ‘Mutation-order 2’). At low dimensionality (A and C), the new environments rewarded only two tasks; at high dimensionality (B and D), the new environments rewarded 34 tasks. With two-point crossover (A and B), only one region was exchanged when creating hybrids; with multiple-point crossover (C and D), up to 100 regions were exchanged. 97 Reproductive isolation in the mutation-order treatment was more sensitive to gene flow than in the ecological treatment (Figs. 4.2A and 4.2B). At the 0.01 migration rate, for instance, the mean hybrid fitness in the mutation-order treatment was > 0.98, but in the ecological treatment reproductive isolation was almost as strong as without migration. The mutation-order 2 treatment was more sensitive to gene flow than the mutation-order 1 treatment (no reproductive isolation at a migration rate of 0.00001). When the environment rewarded for many tasks (high dimensionality), reproductive isolation was often stronger than when the environment rewarded for only two tasks (low dimensionality) (compare Figs. 4.2A and 4.2B, Figs. 4.2C and 4.2D). This pattern was most evident in the ecological treatment, even at moderately high migration rates; for example, the mean hybrid fitness in the ecological treatment at 0.1 migration was 0.97 under low dimensionality but only 0.74 under high dimensionality. In the mutation-order treatments, however, reproductive isolation under high dimensionality at migration rates > 0 was not always stronger than under low dimensionality, showing again that mutation-order was sensitive to gene flow. When hybrids between the evolved demes were created by recombining up to 100 genetic regions (‘multiple-point crossover’), reproductive isolation in the ecological and mutationorder treatments was stronger (Figs. 4.2C and 4.2D). Note that recombination with multiple crossover points was used only to create hybrids for the calculation of postzygotic isolation; all populations were evolved under two-point crossover recombination. The mean hybrid fitness with multiple-point crossover was significantly lower than that with two-point crossover, dropping 33% and 48% in the ecological treatment for low and high dimensionality (respectively) and 53% and 43% in the mutation-order treatments. The difference in strengths of reproductive isolation between ecological and mutation-order treatments was now smaller 98 than that with two-point crossover. Reproductive isolation in the mutation-order treatment remained more sensitive to gene flow than in the ecological treatment. Reproductive isolation in the genetic drift treatment with little migration was significantly greater than with two-point crossover. Interestingly, reproductive isolation in the genetic drift treatment with 0.00001 migration and high dimensionality was even greater than in the mutation-order 2 treatment. 4.3.2 Genetic divergence The genetic divergence under zero migration was significantly higher than that under 0.01 migration for all treatments (Table 4.1), demonstrating that gene flow between populations had a homogenizing effect. Under 0.01 migration, the mutation-order treatments had little genetic divergence (FST < 0.05), which was significantly lower than the ecological treatments, suggesting that the mutation-order treatments were more sensitive to gene flow than the ecological treatments. Interestingly, the drift treatment under zero migration showed high levels of genetic divergence, as high as the ecological and mutation-order treatments for low dimensionality. Under zero migration, the genetic divergence for each treatment for high dimensionality was significantly higher than those for low dimensionality. Under 0.01 migration, the genetic divergence for the ecological treatment for high dimensionality was significantly higher than that for low dimensionality, but for the mutation-order treatments there was no difference in the genetic divergence between low and high dimensionalities. In agreement with these results, the sequences for each pair of demes for treatments under zero migration did not align as well as those under 0.01 migration (Fig. 4.3). These result suggest that the reason that reproductive isolation was mostly absent under a mutation-order process 99 Table 4.1: Genetic divergence (FST ) between demes for each treatment. Migration rate Dimensionality Treatment 0.0 0.01 a – Drift 0.3136 0.0187b a Low Ecological 0.3173 0.1387c Low Mutation-order 1 0.3096a 0.0186b Low Mutation-order 2 0.3002a 0.0216b High Ecological 0.4187d 0.2680e High Mutation-order 1 0.4221d 0.0304b High Mutation-order 2 0.3758f 0.0199b Note: Shared superscript letters indicate that those values are not significantly different (95% bootstrap confidence interval). with gene flow is that the key mutations that allowed organisms to perform tasks were the same (i.e., no genetic divergence for task-related mutations). 4.3.3 Hybrid phenotypes In the ecological treatment, I found that each replicate had, on average, 268.3 hybrids (218.0– 315.8, 95% bootstrap mean C.I.) of 1,000 that contained at least one DMI between their parental components. In the mutation-order treatments, this quantity was 77.1 (45.8–111.2) and 128 (83.5–176.5) of 1,000. Therefore, populations that adapted to different environments accumulated more DMIs than populations that adapted to similar environments. Because hybrids, on average, inherit half the genome of each parent, I expected that hybrids, on average, would inherit half the tasks from each parent (here I focused on the treatments without migration, low dimensionality, and two-point crossover). In the ecological treatment, I found that hybrids were more likely to perform zero, one, or two tasks from one parent and none from the other (Figure 4.4A). Less than 10% of hybrids were able to perform all four tasks. For the mutation-order treatments, most hybrids could perform 100 Mutation-order 1 (no migration) 1 jnjaokkaukjccfsuyypulclkbbycobqfyiy oyjcukoaaklcdclayyponccjbbxmtbmiczx 2 lnjdcliaakjccxyuqyoukucybbjcwocicma yntbbmlaakjccxtryypukccyuoyitcqycia 3 jnjacukaycjccctuyypulccybttctbmrcpc ynjamklaakmccnyuyypupcjybuycuccryzc 4 jnjdcckaaojccoluycuypoynmbrcnbcscic jnjayntcakpccbtuyquyyccnbcoctdcscic 5 lnjachoaycjccbyuyypukccobbrcobmnysa ynycmnmaajaqcjuuymnuncpnbyzctckicmc Mutation-order 1 (0.01 migration) 1 jocsuypuabnmcucybuy jacsuypuabnmcucybuy 2 uaytuspuabnububybcp umytuspuabnububybcp 3 lxcluyurraojjdbuybo uacluyurraojjdbuybo 4 ycctuypuamlmcpcybbt ycctuypuamlmcpcybjt 5 boctrmnuabjmquyyboc boctrmnuabjmquyyboc Figure 4.3: Consensus sequences of the first five evolved replicate pairs of demes in the mutation-order 1 treatment under zero migration (top) and 0.01 migration (bottom). Similar results were observed for the mutation-order 2 treatment. Sequences were 200 instructions in length, but only the loci that differed among each set of five replicates are shown. Derived alleles involved in performing a task are highlighted (black highlight = task 1, gray highlight = task 2, bold font = both). 101 both tasks (Figure 4.4B and 4.4C), but because the parents could also perform both tasks, this information alone did not tell whether hybrids inherited one task from each parent or some other combination. When I analyzed those hybrids that could perform both tasks, I found that most inherited both tasks from just one parent (Figure 4.5), although for mutation-order 2 the difference between those that performed one task from each parent and those that performed both tasks from one parent was not significant. Surprisingly, for the mutation-order 1 treatment there were many hybrids that were fit even though their parental components could perform no tasks or just one task (Figure 4.5A). 4.4 Discussion In this study, I used experimental evolution of digital organisms to compare the strength of postzygotic reproductive isolation generated by ecological and mutation-order processes. I assessed the strength of postzygotic isolation by measuring the mean hybrid fitness relative to each parent in its native environment. I found that, using a two-point crossover recombination method, the mean hybrid fitness was around 55% under ecological divergence but around 83% under a mutation-order process. Other studies have also found that the mean hybrid fitness is lower under divergent selection than under parallel selection. Dettman et al. (2007) found that the mean relative fitness of hybrids between yeast populations evolved in different environments (high-salinity and low-glucose) was around 87%, but hybrids from populations evolved under the same environmental conditions were as fit as their parents. Similar patterns were found in a filamentous fungus by Dettman et al. (2008), although in one of the parental environments hybrids between populations under divergent selection performed better than hybrids under parallel selection. Along with these studies, my study 102 Number of hybrids A 500 Ecological b ab a 400 300 c d 200 c 100 0 0−0 1−0 1−1 2−0 2−1 2−2 Number of tasks performed by hybrid in each environment Number of hybrids B 1000 C Mutation-order 1 1000 800 800 600 Mutation-order 2 600 400 b b a c 400 b c 200 200 0 0 None 1 2 1 and 2 a a None 1 2 1 and 2 Task(s) performed by hybrid Figure 4.4: Number of hybrids able to perform certain tasks (see Methods) for the (A) ecological, (B) mutation-order-1, and (C) mutation-order-2 treatments under zero migration, low dimensionality, and two-point crossover. Each point is a hybrid count (out of 1,000) for a single replicate. Counts from the same replicate are connected by gray lines. The mean hybrid count per category among replicates is indicated by a horizontal bar. Non-significant differences between the mean hybrid counts share the same letter above (or below) the points in each category. 103 500 c A a 400 300 b ab 200 Number of fit hybrids 100 0 0,0 500 1,0 1,1 2,* c B b c 1,0 1,1 400 300 200 100 a 0 0,0 2,* Number of tasks performed by each parental component Figure 4.5: Number of fit hybrids (i.e., can perform two tasks) whose parental components can perform certain tasks (see Methods) for the (A) mutation-order 1 and (B) mutation-order 2 treatments under zero migration, low dimensionality, and two-point crossover. Each point is a hybrid count for a single replicate. Counts from the same replicate are connected by gray lines. The mean hybrid count per category among replicates is indicated by a horizontal bar. Non-significant differences between the mean hybrid counts share the same letter above the points in each category. 104 supports the view that ecological divergence causes stronger reproductive isolation than a mutation-order process. It has been suggested that gene flow during speciation may be common (Coyne and Orr 2004, p. 112; Nosil 2008), which requires that genetic divergence with gene flow be possible. I found that a migration rate of 1% was not enough to prevent genetic divergence under an ecological process (Table 4.1). This finding supports the notion that it is possible for populations under divergent selection in the face of gene flow to continue to diverge. Under a mutation-order process, however, a migration rate of 1% was enough to prevent genetic divergence, which suggests that mutation-order speciation is more sensitive to gene flow than ecological speciation. Nosil and Flaxman (2011) also found in their computer simulations that genetic divergence under a mutation-order process did not occur > 1% gene flow. One of the mutation-order treatments under high dimensionality was even sensitive to a migration rate of 0.00001. I speculate that in this treatment (corresponding to environment 2H), there was one or more large-effect adaptive mutation(s) that, when migrated to the other deme, spread quickly and homogenized the demes. I conclude that different populations under parallel selective pressures probably require almost complete isolation for divergence to occur. Reproductive isolation between populations evolving in high-dimensional environments has been predicted and observed to be stronger than in single or low dimensional environments (Rice and Hostert, 1993; Nosil et al., 2009; Nosil and Harmon, 2009). In Timema walking-stick insects, for example, reproductive isolation showed a positive correlation with environmental dimensionality (Nosil and Sandoval, 2008; Nosil and Harmon, 2009); further examples are reviewed in Nosil et al. (2009). Most empirical studies, however, rely on incomplete measures of dimensionality (imagine the difficulty in accounting for all selective 105 pressures in the field). In this study, I was able to control precisely the number of selective pressures for the low and high dimensionality treatments. I found that under an ecological process, reproductive isolation was stronger between populations in high-dimensional environments than in low dimensional environments. Under a mutation-order process, however, this pattern held only when no migration occurred between populations, but when gene flow was allowed this pattern went away. Our results support previous findings that dimensionality matters for ecological speciation but suggests that for mutation-order speciation with gene flow, environmental dimensionality may not be as important. This conclusion was also supported by my measurements of genetic divergence: there was no difference in genetic divergence between low and high dimensionality for the mutationorder treatments under some gene flow. For the ecological treatment, however, the genetic divergence in high dimensionality was higher than in low dimensionality and higher than the mutation-order treatments, again showing that mutation-order treatments were more sensitive to gene flow. Interestingly, under zero migration the drift treatment (where mutations fixed neutrally) was as high as the ecological and mutation-order treatments under low dimensionality, suggesting that most of the divergence in the ecological and mutationorder treatments was actually the result of neutral fixations and few adaptive mutations. Indeed, in post-hoc analyses I found that about 90% of mutational differences between these treatments were due to neutral fixations, not adaptive mutations. Another result to note is that under zero migration, the ecological and mutation-order treatments had about the same level of genetic divergence, which is closer to my result with multiple-point crossover than two-point crossover, suggesting that in some cases the amount of postzygotic reproductive isolation and genetic divergence are decoupled. 106 This decoupling between reproductive isolation and genetic divergence has been observed in biological populations (Stelkens and Seehausen, 2009; Mac´ Garcia et al., 2012). In these ıas studies, genetic divergence was not found to be a good predictor of sexual dimorphism or assortative mating (Stelkens and Seehausen, 2009; Mac´ Garcia et al., 2012). In some cases, ıas genetically closely related species were ecologically and phenotypically divergent; in other cases, genetically distant species were phenotypically and ecologically close (Stelkens and Seehausen, 2009). One proposed reason for this decoupling is that temporal changes in selection pressures alter the way in which natural and sexual selection interact (Mac´ Garcia ıas et al., 2012). Although assortative mating was not present in my digital populations—there was no mechanism for mate choice—reproductive isolation could not be predicted solely based on genetic divergence. I speculate that the reason was due to the degree of incompatibility between alleles for the different modes of speciation: alleles between populations were not as incompatible under a mutation-order process than under an ecological process. Our results support the notion that reproductive isolation is not directly caused by genetic divergence but is a byproduct of processes that also affect genetic divergence (Pereira et al., 2011). Therefore, in order to determine reproductive isolation between populations, one cannot rely solely on their genetic divergence; reproductive isolation should be measured directly. Traits that are physically modular are hardly broken apart by recombination, hiding genetic incompatibilities that may have formed between populations. To determine whether genetic incompatibilities had formed between my populations but were hidden by the modularity of traits, I re-created hybrids through time using multiple-point crossover recombination rather than two-point crossover recombination. In multiple-point crossover recombination, each locus of a hybrid’s genome had an equal probability of coming from either 107 parent; in this way, modular traits could be broken apart by recombination. I found that the strength of reproductive isolation decreased for both ecological and mutation-order speciation, such that mutation-order speciation was almost as strong as ecological speciation (Figure 4.2). I even see some reproductive isolation in the drift treatment, showing that incompatibilities also formed, but at a much slower rate than speciation by natural selection. These findings show that genetic incompatibilities were hidden by the modularity of traits. In other words, genetic incompatibilities that formed between populations were not always seen in hybrids because two-point crossover recombination did not break apart co-adapted gene complexes coding for a task. I note that the genetic architecture of my populations evolved under two-point crossover recombination, not multiple-point crossover recombination, and thus the modularity of traits and formation of genetic incompatibilities may be different under a different recombination method. Part of the reason that hybrids were more unfit in the ecological treatment than in the mutation-order treatment was that in the ecological treatment more genetic incompatibilities (DMIs) formed between populations. This result supports the view that genetic incompatibilities are an important cause of ecological speciation (Rundle and Nosil, 2005). Another reason that hybrids were more unfit in the ecological treatment was that for a hybrid to be fully fit it had to inherit both sets of tasks from both parents (i.e., four tasks), whereas for the mutation-order treatment, hybrids required only two tasks to be fit. In the ecological treatment, most hybrids inherited either one one two tasks from one parent and none from the other, and in the mutation-order treatments, most hybrids inherited both tasks from one parent, although in the second mutation-order treatment hybrids often inherited one task from each parent. 108 I found that the selection coefficients of adaptive alleles in the low dimensionality environments were higher than that in the high dimensionality environments. An opposite trend may have made it difficult to know whether it was higher dimensionality or stronger selection that resulted in hybrids being less fit under high dimensionality than low dimensionality. The smaller selection coefficients in the high dimensionality environments may seem puzzling at first. But given that each additional task an organism could perform gives it an equal amount of merit, the higher the merit, the less an additional task contributes to the total merit. Therefore, the more tasks organisms can perform in the high dimensionality environment, the less beneficial each one becomes (i.e., diminishing returns). The strengths of selection in either environment are nevertheless high overall, but it is not uncommon for selection to be high in new environments (e.g., Lenski et al., 1991; Dettman et al., 2007). Future studies could investigate how the strength of selection may affect the strength of postzygotic isolation by manipulating the selection coefficients in each environment. In summary, I used the artificial life platform Avida, which allowed us to precisely control the type of selection (divergent or uniform), to compare the strength of reproductive isolation between ecological and mutation-order speciation. By accurately measuring the fitness of hybrids between populations, I showed that ecological speciation formed stronger postzygotic isolation than mutation-order speciation, although they were not so different when recombination involved crossover at multiple points. In addition, Avida allowed us to test various specific migration rates during the evolution of population pairs, where I found that mutation-order speciation was more sensitive to gene flow than ecological speciation. I was also able to control the number of selection pressures in each population, and I found that environments with high dimensionality formed stronger reproductive isolation than those with low dimensionality. These results support ideas brought up in the literature 109 but which have been difficult to test in biological organisms. Avida provided a platform for us to test these ideas much more easily, and although digital organisms are more simplistic than biological organisms, they are a genetic system that evolves and speciates and therefore allows us to test the generality of hypotheses about speciation, which often do not require the specific details about how biological organisms work. Acknowledgments I would like to thank my co-author on this work, Luke Harmon, who reviewed this manuscript several times and provided me with great ideas. I would also like to thank B. L. Williams and the Digital Evolution Laboratory at MSU for discussion and suggestions for improvement. I am grateful to S. Singhal and two anonymous reviewers for their comments that helped improve the quality of the manuscript. This material is based in part upon work supported by the National Science Foundation under Cooperative Agreement No. DBI-0939454. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the National Science Foundation. 110 LITERATURE CITED 111 LITERATURE CITED Adami, C. (2006). Digital genetics: unraveling the genetic basis of evolution. Nature Rev. Genet., 7:109–118. Burton, R., Rawson, P. D., and Edmands, S. (1999). Genetic architecture of physiological phenotypes: empirical evidence for coadapted gene complexes. Amer. Zool., 39:451–462. Chow, S. S., Wilke, C. O., Ofria, C., Lenski, R. E., and Adami, C. (2004). Adaptive radiation from resource competition in digital organisms. Science, 305:84–86. Conner, J. K. and Hartl, D. L. (2004). A Primer of Ecological Genetics. Sinauer Associates, Inc., Sunderland, MA. Coyne, J. A. and Orr, H. A. (2004). Speciation. Sinauer Associates, Inc., Sunderland, MA. Dettman, J. R., Anderson, J. B., and Kohn, L. M. (2008). Divergent adaptation promotes reproductive isolation among experimental populations of the filamentous fungus Neurospora. BMC Evol. Biol., 8:35. Dettman, J. R., Sirjusingh, C., Kohn, L. M., and Anderson, J. B. (2007). Incipient speciation by divergent adaptation and antagonistic epistasis in yeast. Nature, 447:585–588. Elena, S. F., Wilke, C. O., Ofria, C., and Lenski, R. E. (2007). Effects of population size and mutation rate on the evolution of mutational robustness. Evolution, 61:666–674. Fry, J. D. (2009). Laboratory experiments on speciation, pages 631–656. University of California Press, Berkeley, CA. Garant, D., Ford, S. E., and Hendry, A. P. (2006). The multifarious effects of dispersal and gene flow on contemporary adaptation. Funct. Ecol., 21:434–443. Gillespie, J. H. (2004). Population genetics: a concise guide. Johns Hopkins University Press, Baltimore, MD. Hartl, D. L. and Clark, A. G. (1997). Principles of population genetics. Sinauer Associaties, Inc., Sunderland, MA. Lenski, R. E., Ofria, C., Collier, T. C., and Adami, C. (1999). Genome complexity, robustness, and genetic interactions in digital organisms. Nature, 400:661–664. Lenski, R. E., Ofria, C., Pennock, R. T., and Adami, C. (2003). The evolutionary origin of complex features. Nature, 423:139–144. Lenski, R. E., Rose, M. R., Simpson, S. C., and Tadler, S. C. (1991). Long-term experimental evolution in Escherichia coli. i. adaptation and divergence during 2,000 generations. American Naturalist, 138:1351–1341. 112 Li, Z., Pinson, S. R. M., Paterson, A. H., Park, W. D., and Stansel, J. W. (1997). Genetics of hybrid sterility and hybrid breakdown in an intersubspecific rice (Oryza sativa l.) population. Genetics, 145:1139–1148. Mac´ Garcia, C., Smith, G., Gonz´lez Zuarth, C., Graves, J. A., and Ritchie, M. G. (2012). ıas a Variation in sexual dimorphism and assortative mating do not predict genetic divergence in the sexually dimorphic goodeid fish Girardinichthys multiradiatus. Curr. Zool., 58:440– 452. Misevic, D., Ofria, C., and Lenski, R. E. (2010). Experiments with digital organisms on the origin and maintenance of sex in changing environments. J. Hered., 101:S46–S54. Nosil, P. (2008). Speciation with gene flow could be common. Mol. Ecol., 17:2103–2106. Nosil, P. and Flaxman, S. M. (2011). Conditions for mutation-order speciation. Proc. R. Soc. B, 278:399–407. Nosil, P. and Harmon, L. J. (2009). Niche dimensionality and ecological speciation. In K., B. R., J., B., and D., S., editors, Speciation and patterns of diversity, pages 127–154. Cambridge University Press. Nosil, P., Harmon, L. J., and Seehausen, O. (2009). Ecological explanations for (incomplete) speciation. Trends Ecol. Evol., 24:145–156. Nosil, P. and Sandoval, C. P. (2008). Ecological nich dimensionality and the evolutionary diversification of stick insects. PLoS ONE, 3:e1907. Ofria, C. and Wilke, C. O. (2004). Avida: a software platform for research in computational evolutionary biology. Artif. Life, 10:191–229. Ostrowski, E. A., Ofria, C., and Lenski, R. E. (1997). Ecological specialization and adaptive decay in digital organisms. American Naturalist, 169:E1–E20. Pereira, R. J., Monahan, W. B., and Wake, D. B. (2011). Predictors for reproductive isolation in a ring species complex following genetic and ecological divergence. BMC Evol. Biol., 11:194. R¨s¨nen, K. and Hendry, A. P. (2008). Disentangling interactions between adaptive diveraa gence and gene flow when ecology drives diversification. Ecol. Lett., 11:624–636. Rice, W. R. and Hostert, E. E. (1993). Laboratory experiments in speciation: what have we learned in 40 years? Evolution, 47:1637–1653. Rundle, H. D. and Nosil, P. (2005). Ecological speciation. Ecol. Lett., 8:336–352. Schluter, D. (2009). Evidence for ecological speciation and its alternative. Science, 323:737– 741. Schluter, D. and Conte, G. L. (2009). Genetics and ecological speciation. Proc. Natl. Acad. Sci. USA, 106:9955–9962. 113 Sobel, J. M., Chen, G. F., Watt, L. R., and Schemske, D. W. (2010). The biology of speciation. Evolution, 64:295–315. Stelkens, R. B. and Seehausen, O. (2009). Phenotypic divergence but not genetic distance predicts assortative mating among species of a cichlid fish radiation. J. Evol. Biol., 22:16791694. Wilke, C. O. and Adami, C. (2002). The biology of digital organisms. Trends Ecol. Evol., 17:528–532. 114 Chapter 5 Experimental evolution of the snowball effect shows the importance of complex epistasis in speciation 5.1 Introduction Biological speciation is the evolution of barriers that hinder interbreeding between populations (‘reproductive isolating barriers’) (Coyne and Orr, 2004). Postzygotic barriers cause hybrids between species to be sterile or inviable, sometimes because genes from different species are incompatible (Coyne and Orr, 2004). The Dobzhansky-Muller model of postzygotic isolation (Fig. 5.1) proposes that genetic incompatibilities form as a byproduct of the genetic divergence between independently evolving populations (Orr, 1995). The Dobzhansky-Muller model predicts that the number of genetic incompatibilities, called ‘DobzhanskyMuller incompatibilities’ (DMIs), should increase faster than linearly through time (Orr, 1995; Orr and Turelli, 2001). For example, pairwise DMIs (i.e., DMIs between two alleles) should increase quadratically through time (Orr and Turelli, 2001). This phenomenon has been called the ‘snowball effect’ (Coyne and Orr, 2004). 115 e E d D c C B b a A abcde Figure 5.1: The Dobzhansky-Muller model of postzygotic isolation. A population with haploid genotype abcde (bottom) becomes geographically divided into two, and each population independently evolves through time (thick arrows, time progresses upward). In the first population, allele a is substituted with allele A, while in the second population allele b is substituted with allele B. At this point, if the populations were to come into contact and hybridize, their hybrids would have genotypes Abcde, aBcde, or ABcde. Because alleles A and B evolved independently and thus may only function properly in the background in which they evolved, hybrids with genotype ABcde may be sterile or inviable. In this case, alleles A and B are said to form a Dobzhansky-Muller incompatibility (DMI) (arrow from B to A). The populations could have instead hybridized after two or more alleles have fixed within each lineage (e.g., alleles A, C, and E in the first population and alleles B and D in the second), which would have resulted in hybrids with many more possible DMIs. DMIs between derived alleles (‘derived-derived’ DMIs) are thin solid arrows, and DMIs between derived and ancestral alleles (‘derived-ancestral’ DMIs) are thin dashed arrows. (Modified from (Orr, 1995).) 116 Several tests of the snowball effect have not supported it (Lijtmaer et al., 2003; Mendelson et al., 2004; Gourbi`re and Mallet, 2010), concluding that there is a ‘missing snowball.’ e These tests, however, relied on indirect methods because testing the snowball effect directly is difficult (Mendelson et al., 2004). Rather than using a single pair of species and following it through time, researchers have used different pairs of species diverged at different times (Malon and Fontenot, 2008; Scopece et al., 2008). For the number of DMIs, the strength of postzygotic isolation, often measured as the reduction in hybrid viability or fertility, has been used. However, for the strength of postzygotic isolation to be a good proxy for the number of DMIs, the fitness effects of DMIs on hybrid fitness must be additive (Mendelson et al., 2004; Bolnick and Near, 2005) (e.g., twice the number of DMIs should result in twice the isolation), but whether this is true is not known (Bolnick and Near, 2005; Presgraves, 2010). Better estimates of the number of DMIs between two species were recently carried out by two studies (Matute et al., 2010; Moyle and Makazato, 2010), both supporting the snowball effect. To estimate the number of DMIs, they introgressed single genetic regions of one species into the genetic background of the other and counted the number of introgressions with reduced viability or fertility. However, this method cannot identify individual DMIs but rather whether a genetic region of one species is incompatible with something in the other. As with the previous methods, they relied on the ages of species pairs rather than following a single species pair through time. In addition, because these studies could only obtain at most three species pairs with which to count the number of DMIs, their quadratic fit to the data was not statistically powerful. To provide an example of what is required to identify a DMI, suppose that allele D in Fig. 5.1 has just fixed in one population, and I want to test whether it is incompatible 117 with allele C in the other population. One option is to isolate these alleles in the ancestral genetic background (abcde) to construct the genotype abCDe and measure its fitness. In this constructed genotype, however, alleles D and b as well as alleles C and a may also be incompatible (Fig. 5.1), and therefore confound the effect of C and D together. Every other possible construction of a genotype that includes C and D (i.e., aBCDe, AbCDe, and ABCDe) also contains other possibly confounding incompatibilities. To prevent these confounding factors, one may only consider alleles that were compatible with the ancestral background. In the example above, this means that one must first verify that genotypes abCde and abcDe had normal fitness before testing abCDe. In this study, I conducted experimental evolution to test whether pairwise DMIs increased quadratically through time between evolving populations of digital organisms. I also measured hybrid inviability through time in order to determine whether there was a missing snowball. Finally, I determined the ratio of derived-derived and derived-ancestral DMIs and the probability than a pairwise interaction forms a DMI, which in the literature has been assumed to be constant. I used Avida (Ofria and Wilke, 2004) (see p. 6), an artificial life research platform, to conduct my experiments with digital organisms. I used digital organisms for several reasons: I can observe thousands of generations in minutes, easily manipulate genomes, which allowed us to develop a method to accurately count the number of individual pairwise DMIs as they arose, and have an accurate historical record. 5.2 Results To generate the ancestral genotype, I first founded a population with an organism that could replicate sexually but could not perform any tasks. I then let this population evolve for about 118 1.5 million generations, and I took the genotype of the most common organism as the ancestral genotype. Using this ancestor, I then founded 40 independent populations and allowed them to evolve independently for about 10,000 generations. The configuration parameters, including the environmental conditions, were the same as those of the ancestor. Every 400 generations, the entire population was saved for later analysis. Because the ancestor was well-adapted to its environment, the genetic divergence between replicate populations was mostly due to fixation of neutral mutations. Thus, the rate of substitution for each population was approximately constant, satisfying an assumption of the snowball effect when it is analyzed through time (Orr, 1995). The mean relative fitness of hybrids between pairs of populations at the end of the runs was 0.91 (range: 0.68–0.97), indicating that incomplete postzygotic isolation had evolved. From now on, I refer to these populations as ‘species.’ 5.2.1 Pairwise DMIs increased quadratically through time To count the number of pairwise DMIs between two species, I separately counted the number of derived-ancestral and derived-derived DMIs (Fig. 5.2). To find derived-ancestral DMIs, I first searched for single derived alleles of each species that were incompatible with the ancestral background (e.g., allele E in Fig. 5.2A, step 1). I defined an allele as incompatible with the ancestral background if the fitness of the ancestral genotype with that allele alone was < 0.75 relative to the original ancestor. To determine whether the incompatibility was due to a single ancestral allele (thereby forming a derived-ancestral DMI), I searched for another derived allele of the same species that rescued the incompatibility (e.g., allele C in Fig. 5.2A, steps 2 and 3). I defined an incompatibly as rescued by another allele if the relative fitness of the ancestral genotype with both alleles was > 0.99. To ensure that the rescue allele was itself not involved in other DMIs, I verified that its fitness in the ancestral 119 background was also > 0.99. I excluded testing rescue alleles that appeared after the derived allele because in a derived-ancestral DMI a derived allele cannot be incompatible with any current ancestral alleles (e.g., in Fig. 5.1, allele C cannot be incompatible with e). To find derived-derived DMIs, I searched for single derived alleles from both species that were each compatible with the ancestral background but together were incompatible (Fig. 5.2B). Using this two-part method, I counted the total number of pairwise DMIs for each of the 20 replicate pairs of populations every 400 generations. To determine whether a linear (ax) or a quadratic (ax2 + bx) model best described the accumulation of pairwise DMIs through time, I fit these two models to the whole dataset (n = 520) and to each replicate individually (each n = 26). Because there are no pairwise DMIs at the moment of geographic isolation, the models do not have a constant term (i.e., the intercept is 0). I estimated the parameters of each model using maximum likelihood with a Gamma distribution for DMIs and compared the models using AIC (Bolker, 2008). I found that the quadratic model explained the whole dataset better than the linear model (Fig. 5.3A), although there was considerable variation per replicate. These results indicate that the overall accumulation of pairwise DMIs was consistent with the snowball effect. 5.2.2 Hybrid inviability increased linearly through time To measure hybrid inviability (i.e., number of unfit hybrids) through time, I first created 10,000 hybrids between each replicate pair of species every 400 generations. Hybrids were created in the same way Avida recombines two genotypes: two random but homologous regions of the parental genomes were exchanged. Note that hybrids were created after the experiments were done, using the populations I saved every 400 generations; no hybridization between populations occurred during their evolution. I then measured the fitness of each 120 A A b C d E B a B c 1 D e 1 A b c d e a B c d e a b C d e a b c D e a b c d E 2 A A b c d E a b C d b c D e a 2 b C D e E 3 Figure 5.2: Illustration of the method for identifying (A) derived-ancestral and (B) derivedderived pairwise DMIs (see text). (A) Step 1: I tested the individual fitness effect of each derived allele (black bars) of a species on the ancestral background. Step 2: For each derived allele that was incompatible with the ancestral background (indicated by the black triangle pointing down), I tested the individual effect of the other derived alleles on that genetic background. Step 3: If a second derived allele resulted in high fitness (indicated by the gray triangle pointing up), the ancestral allele at that second locus was incompatible with the original derived allele. (B) Step 1: I tested the individual fitness effect of each derived allele (gray bars) of a species on the ancestral background. Step 2: For each derived allele that was compatible with the ancestral background, I tested the individual effect of each derived allele (itself compatible with the ancestral background) of the other species. Step 3: If the second allele lowered the fitness, then these two allele must be incompatible. 121 A Number of DMIs 15 10 5 0 0 Number of unfit hybrids 1,500 2,500 5,000 7,500 10,000 2,500 5,000 7,500 10,000 B 1,000 500 0 0 Time (generations) Figure 5.3: (A) Mean number of pairwise DMIs through time. Each point represents the mean of 20 replicate runs. The black curve represents the quadratic model of the data with parameters estimated using maximum likelihood. The gray lines represent the bootstrap 95% confidence intervals of the means. (B) Mean number of unfit hybrids (out of 10,000) between populations through time. Each point represents the mean of 20 replicate runs. The black line represents the linear model of the data with parameters estimated using maximum likelihood. The gray lines represent the bootstrap 95% confidence intervals of the means. 122 hybrid and counted the number (out of 10,000) that had a relative fitness < 0.75. I found that the overall number of unfit hybrids through time increased linearly (Fig. 5.3B), although there was considerable variation per replicate. These results are consistent with a missing snowball for hybrid inviability and therefore suggest that the fitness effects of DMIs were not additive. 5.2.3 Secondary alleles rescue pairwise DMIs Previous studies have found a linear increase in hybrid inviability through time, but conclusions that these results indicate the absence of a snowball effect require that DMIs be additive. If pairwise DMIs were additive, I would expect that hybrids carrying at least one pairwise DMI be unfit because a single pairwise DMI should reduce the fitness of the carrier to < 0.75. I found, however, that not all hybrids carrying at least one pairwise DMI were unfit (Fig. 5.4), indicating that pairwise DMIs were not additive. One possible reason pairwise DMIs were not additive is that other derived alleles present in the hybrid rescued pairwise DMIs (i.e., the true incompatibility was greater than pairwise). This hypothesis predicts that fit hybrids carrying a pairwise DMI are more likely to carry a rescue allele than unfit hybrids. To test this prediction, I first identified all possible rescue alleles for all known pairwise DMIs by searching for single derived alleles from either species that rescued each known pairwise DMI. Then, for hybrids carrying a pairwise DMI (at generation 10,000), I calculated the proportion that also carried a rescue allele. I found that 97% of fit hybrids carried a rescue allele compared to only 45% of unfit hybrids. This finding suggests that certain alleles rescued pairwise DMIs, and this complex interaction could explain why pairwise DMIs were not additive and therefore why the mean hybrid inviability increased only linearly. 123 3,000 All hybrids w/ pwDMIs Unfit hybrids w/ pwDMIs Number of hybrids 2,500 2,000 1,500 1,000 500 0 0 2,500 5,000 7,500 10,000 Time (generations) Figure 5.4: Mean number of all hybrids ( ) and unfit hybrids ( ) (out of 10,000) with at least one pairwise DMI through time. Each point represents the mean of 20 replicate runs. The gray lines represent the bootstrap 95% confidence intervals of the means. 5.2.4 Derived-ancestral DMIs occur as often as derived-derived Derived alleles have been predicted to be three times more likely than ancestral alleles to be involved in pairwise DMIs (Orr, 1995). To test this prediction, I counted the number of times a derived allele appeared in a DMI (once in a derived-ancestral DMI and twice in a derived-derived DMI) and the number of times an ancestral allele appeared in a DMI (once in a derived-ancestral DMI). I made these counts at 10,000 generations for each pair of populations, and I calculated the mean of the ratios between the number of derived alleles and ancestral alleles found in all DMIs. I found that derived alleles are 3.06 (2.41–3.79, 95% bootstrap C.I.) times more likely than ancestral alleles to appear in pairwise DMIs, which experimentally supports the prediction. However, the number of derived-ancestral and derived-derived DMIs were comparable, suggesting that hybrid inviability due to a 124 missing dependent allele in the same species was just as likely due to incompatible derived alleles between species. 5.3 Discussion Although the mean accumulation of pairwise DMIs increased quadratically, there was considerable variation in the pattern of accumulation for individual species pairs. There are at least three main possibilities for this variation. First, the probability p that two alleles are incompatible may not be constant through time. For example, a new derived allele may form multiple pairwise incompatibilities with ancestral or derived alleles of the other population, increasing the value of p temporarily. Second, as an evolving population navigates its neutral landscape, a derived allele that once caused a DMI may later be replaced by an allele that does not cause any DMIs. Third, I used the majority-rule consensus sequence of a population for my analyses, but alleles in a consensus sequence are not necessarily fixed. For example, an allele that was present in 51% of the population would appear in the consensus sequence, but if that allele later decreases in frequency through drift it could disappear from the consensus sequence. If such an allele was involved in DMIs, the estimated number of DMIs would change over time. All of my population pairs underwent evolution in the same environmental conditions, yet showed variation in the accumulation of DMIs. This variation suggests that in natural populations, where environmental conditions between any taxa are rarely identical, the accumulation of DMIs should also vary considerably. Alleles that rescue hybrid fitness are not unique to digital organisms. In Drosophila, several ‘hybrid rescue mutations’ recover the viability or fertility of hybrids between D. melanogaster and closely-related species. Note that the Drosophila rescue alleles were ar- 125 tificially selected mutations, not derived alleles that were fixed in natural populations, so the importance of rescue alleles in the wild is currently unknown. Two hypotheses explain how rescue alleles may interact with DMIs. In the first hypothesis, rescue alleles “suppress the effects of the loci causing hybrid problems” (Coyne and Orr, 2004). In this case, rescue alleles may be products of genetic redundancy, which itself may have evolved as a ‘buffering mechanism’ against deleterious mutations (Wagner, 1999; Elena et al., 2006). For example, a genotype that gains a new allele performing overlapping functions with another allele at a different locus will be more robust to deleterious mutations that affect those functions. Although I do not know whether rescue alleles in my system are redundant, studies have shown that sexual, complex digital organisms with high mutation rates exhibit mutational robustness (Lenski et al., 1999; Wilke et al., 2001; Misevic et al., 2006), which can evolve by genetic redundancy (Elena et al., 2007). According to the second hypothesis, rescue alleles “represent mutations at the actual loci that cause the death or sterility of hybrids” (Coyne and Orr, 2004). This hypothesis implies that hybrid incompatibilities thought to involve only two alleles actually involve three or more, as in the case with the hybrid rescue mutation Hmr in Drosophila (Barbash et al., 2000; Orr and Irving, 2000). Similarly, if this hypothesis applies to the rescue alleles I discovered in digital organisms, then any presumed ‘pairwise’ DMI that was rescued by a third allele was, in fact, a three-way DMI. Therefore, rescue alleles may provide evidence that complex DMIs, which are exceptionally difficult to identify in biological organisms, are common in biological and digital organisms. In summary, using an artificial life software I found that pairwise DMIs accumulate quadratically through time, supporting the snowball effect and the Dobzhansky-Muller model of postzygotic isolation. In addition, I discovered that the number of derived-ancestral and derived-derived DMIs are similar, suggesting that hybrid inviability due to a missing allele 126 in the same species were just as likely due to incompatible derived alleles between species. When I used the strength of postzygotic isolation as a proxy for the number of DMIs, I found a linear, rather than quadratic, relationship with divergence time. This discrepancy was at least partially caused by rescue alleles, which I found recovered the negative effects of DMIs in hybrids, disrupting the pattern of quadratic increase of DMIs. Our findings indicate that pairwise DMIs are insufficient to account for the complexity of epistatic interactions among alleles within and between species. Thus, my results highlight the importance of complex interactions in the genetics of reproductive isolation. 5.4 5.4.1 Methods Experimental configuration The environment was configured to reward the nine default tasks and 68 additional threeinput tasks. All tasks were set to provide a resource value of 1 in an additive fashion; resources were unlimited. The maximum population size was set to 100 organisms, and the length of each organism’s genome was fixed at 500 instructions. The ‘copy’ mutation probability—the probability that an organism would copy a random instruction rather than its own to its offspring—was set to 0.1 per genome per generation (all other mutation probabilities were set to 0.0). Mutation to the ‘h-copy’ instruction was turned off to prevent organisms from using up their genomic space with h-copy instructions (rather than task-related instructions), which was a common way to improve their replication efficiency in preliminary runs. Offspring were configured to replace a random organism in the population. 127 5.4.2 Statistics Statistical analyses were carried out using R (ver. 2.15.1) (R Core Team, 2013). To perform the AIC analyses, I used the bbmle package (Bolker, 2008) and the mle2 function with the ‘Nelder-Mead’ optimization method. 5.5 Acknowledgments I thank A. P. Dreyer, E. M. Swanson, R. E. Lenski, J. W. Boughman, E. Weigel, E. Dittmar, L. J. Harmon, and D. R. Matute for helpful comments on the manuscript. This material is based in part upon work supported by the National Science Foundation under Cooperative Agreement No. DBI-0939454. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author and do not necessarily reflect the views of the National Science Foundation. 128 LITERATURE CITED 129 LITERATURE CITED Barbash, D. A., Roote, J., and Ashburner, M. (2000). The Drosophila melanogaster Hybrid male rescue gene causes inviability in male and female species hybrids. Genetics, 154:1747– 1771. Bolker, B. M. (2008). Ecological models and data in R. Princeton University Press, Princeton, NJ. Bolnick, D. I. and Near, T. J. (2005). Tempo of hybrid inviability in centrarchid fishes (teleostei: Centrarchidae). Evolution, 59:1754–1767. Coyne, J. A. and Orr, H. A. (2004). Speciation. Sinauer Associates, Inc., Sunderland, MA. Elena, S. F., Carrasco, P., Dar`s, J. A., and Sanju´n, R. (2006). Mechanisms of genetic o a robustness in RNA viruses. EMBO Reps., 7:168–173. Elena, S. F., Wilke, C. O., Ofria, C., and Lenski, R. E. (2007). Effects of population size and mutation rate on the evolution of mutational robustness. Evolution, 61:666–674. Gourbi`re, S. and Mallet, J. (2010). Are species real? the shape of the species boundary e with exponential failure, reinforcement, and the “missing snowball”. Evolution, 64:1–24. Lenski, R. E., Ofria, C., Collier, T. C., and Adami, C. (1999). Genome complexity, robustness, and genetic interactions in digital organisms. Nature, 400:661–664. Lijtmaer, D. A., Mahler, B., and Turbaro, P. L. (2003). Hybridization and postzygotic isolation patterns in pigeons and doves. Evolution, 57:1411–1418. Malon, J. H. and Fontenot, B. E. (2008). Patterns of reproductive isolation in toads. PLoS ONE, 3:e3900. Matute, D. R., Butler, I. A., Turissini, D. A., and Coyne, J. A. (2010). A test of the snowball theory for the rate of evolution of hybrid incompatibilities. Science, 329:1518–1521. Mendelson, T. C., Inouye, B. D., and Rausher, M. D. (2004). Quantifying patterns in the evolution of reproductive isolation. Evolution, 58:1424–1433. Misevic, D., Ofria, C., and Lenski, R. E. (2006). Sexual reproduction reshapes the genetic architecture of digital organisms. Proc. R. Soc. B, 273:457–464. Moyle, L. C. and Makazato, T. (2010). Hybrid incompatibility “snowballs” between Solanum species. Science, 329:1521–1523. Ofria, C. and Wilke, C. O. (2004). Avida: a software platform for research in computational evolutionary biology. Artif. Life, 10:191–229. Orr, H. A. (1995). The population genetics of speciation: The evolution of hybrid incompatibilities. Genetics, 139:1805–1813. 130 Orr, H. A. and Irving, S. (2000). Genetic analysis of the Hybrid male rescue locus of Drosophila. Genetics, 155:225–231. Orr, H. A. and Turelli, M. (2001). The evolution of postzygotic isolation: accumulating dobzhansky-muller incompatibilities. Evolution, 55:1085–1094. Presgraves, D. C. (2010). Speciation genetics: search for the missing snowball. Curr. Biol., 20:R1073–R1074. R Core Team (2013). R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. Scopece, G., Widmer, A., and Cozzolino, S. (2008). Evolution of postzygotic reproductive isolation in a guild of deceptive orchids. Am. Nat., 171:315–326. Wagner, A. (1999). Redundant gene functions and natural selection. J. Evol. Biol, 12:1–16. Wilke, C. O., Wang, J. L., Ofria, C., Lenski, R. E., and Adami, C. (2001). Evolution of digital organisms at high mutation rates leads to survival of the flattest. Nature, 412:331–333. 131