<<

HIGHLIGHTED ARTICLE | INVESTIGATION

Evolution of Rates in Rapidly Adapting Asexual Populations

Benjamin H. Good*,†,‡ and Michael M. Desai*,†,‡,1 *Department of Organismic and Evolutionary Biology, †Department of Physics, and ‡Faculty of Arts and Sciences Center for Systems Biology, Harvard University, Cambridge, Massachusetts 02138

ABSTRACT Mutator and antimutator alleles often arise and spread in both natural microbial populations and laboratory evolution experiments. The evolutionary dynamics of these mutation rate modifiers are determined by indirect selection on linked beneficial and deleterious . These indirect selection pressures have been the focus of much earlier theoretical and empirical work, but we still have a limited analytical understanding of how the interplay between hitchhiking and deleterious load influences the fates of modifier alleles. Our understanding is particularly limited when clonal interference is common, which is the regime of primary interest in laboratory microbial evolution experiments. Here, we calculate the fixation probability of a mutator or antimutator allele in a rapidly adapting asexual population, and we show how this quantity depends on the population size, the beneficial and deleterious mutation rates, and the strength of a typical driver mutation. In the absence of deleterious mutations, we find that clonal interference enhances the fixation probability of mutators, even as they provide a diminishing benefit to the overall rate of adaptation. When deleterious mutations are included, pushes the population toward a stable mutation rate that can be suboptimal for the adaptation of the population as a whole. The approach to this stable mutation rate is not necessarily monotonic: even in the absence of epistasis, selection can favor mutator and antimutator alleles that “overshoot” the stable mutation rate by substantial amounts.

KEYWORDS mutator; antimutator; hitchhiking; deleterious load

NA replication occurs with extremely high fidelity, de- higher than the wild type, can be propagated for thousands of Dspite taking place in the noisy environment of the cell. For generations in the laboratory without significant loss of via- example, laboratory strains of Escherichia coli produce a point bility (McDonald et al. 2012; Wiser et al. 2013). This suggests mutation at a rate of roughly one nucleotide per 10 billion that mutation rates are not maintained solely by hard selec- copied (Wielgoss et al. 2011; Lee et al. 2012), which implies tion, but also by more direct evolutionary competition be- that hundreds of generations can elapse before a single mu- tween strains with different mutation rates. These variants tation is introduced into the . To achieve such low feel the effects of natural selection indirectly, by being linked error rates, bacteria and employ a complex array to other mutations that directly influence fitness. of cellular machinery, which must be maintained by natural Strains with higher mutation rates are more likely to be selection. linked to deleterious mutations, and will therefore experience One explanation for the low observed error rates is that the an effective fitness cost. This cost can be measured in head-to- genome encodes a large number of functions, all of which are head competitions between mutator and wild-type strains essential for survival. Low mutation rates could then emerge (Tröbner and Piechocki 1981; Chao and Cox 1983; Chao from hard selection against these lethalerrors. But, in practice, et al. 1983; Giraud et al. 2001; Thompson et al. 2006; observed mutation rates lie far below the levels that would Gentile et al. 2011). In the absence of beneficial mutations, quickly result in extinction. This is evident from the fact that natural selection will therefore act to decrease the mutation mutator strains, whose mutation rates are 10- to 1000-fold rate, until it is eventually balanced by genetic drift, mutation, or other physiological costs. A large body of previous theo- Copyright © 2016 by the Society of America doi: 10.1534/genetics.116.193565 retical work has explored these dynamics (Kimura 1967; Manuscript received July 7, 2016; accepted for publication September 13, 2016; Liberman and Feldman 1986; Dawson 1998, 1999; Johnson published Early Online September 19, 2016. 1Corresponding author: Harvard University, 457.20 Northwest Laboratory, 52 Oxford 1999a; Lynch 2008; Desai and Fisher 2011; Lynch 2011; St., Cambridge, MA 02138. E-mail: [email protected] Soderberg and Berg 2011; James and Jain 2015).

Genetics, Vol. 204, 1249–1266 November 2016 1249 When beneficial mutations are available, lineages with mutation rate by a factor r (with r . 1 corresponding to muta- higher mutation rates are also more likely to produce and tor alleles and r , 1 corresponding to antimutators). We con- hitchhike with a successful beneficial variant. In the absence sider both small changes in mutation rate (r 1), as well as of deleterious mutations, natural selection will therefore act substantial changes (where jlogðrÞj 1). to increase the mutation rate, until the supply of beneficial We find that clonal interference amplifies the beneficial mutations is eventually exhausted. In accordance with these advantage of mutator alleles, even as it constrains the overall expectations, mutator alleles are often found to spontaneously rate of adaptation. For large r this can be a substantial effect, arise and spread in rapidly adapting microbial populations in resulting in .100-fold increases in the probability of fixation. the laboratory (Sniegowski et al. 1997; Notley-McRobb et al. However, clonal interference also amplifies the effects of the 2002; Shaver et al. 2002; Pal et al. 2007; Voordeckers et al. deleterious load, resulting in a much narrower window of 2015), and are often correlated with pathogenic lifestyles in mutator favorability. We show that this window depends the wild (LeClerc et al. 1996; Matic et al. 1997; Oliver et al. most strongly on the strengths of typical driver mutations, 2000; Bjorkholm et al. 2001; Denamur et al. 2002; Giraud et al. and relatively weakly on the rate at which beneficial muta- 2002; Richardson et al. 2002; Prunier et al. 2003; Watson tions arise. With the steady production of such modifier al- et al. 2004; del Campo et al. 2005; Labat et al. 2005). leles, the mutation rate will evolve toward a stable point at In a few cases, laboratory populations that have previously which neither mutator nor antimutator alleles are favored. fixed a mutator allele have been shown to evolve a lower Surprisingly, we find that the approach to this stable point is mutation rate later in the experiment, suggesting that mutation not necessarily monotonic, even in the absence of epistasis. rate evolution is an ongoing process (Tröbner and Piechocki Instead, alleles that overshoot the stable point can be posi- 1984; Notley-McRobb et al. 2002; McDonald et al. 2012; tively selected, and selection must then act to change the Turrientes et al. 2013; Wielgoss et al. 2013). Because these mutation rates in the opposite direction. populations often continue to increase in fitness during this process (Wielgoss et al. 2013), the success of mutator and Model antimutator alleles must depend on the interplay between beneficial and deleterious mutations, rather than the complete Our goal is to understand how beneficial and deleterious cessation of adaptation. However, our understanding of this mutations influence the fates of mutation rate modifiers in interplay remains incomplete. A few simulation studies have rapidly adapting populations. We focus on the simplest pos- shown how these effects can, in principle, favor either in- sible model in which we can address this question. Specifically, creases or decreases in mutation rates, depending on the spe- we consider an asexual haploid population of constant size N cific population parameters (Taddei et al. 1997; Tenaillon et al. (our analysis also applies to diploids when all mutations have 1999, 2000; Travis and Travis 2002). Some analytical progress intermediate dominance effects). We assume beneficial and has also been made in the case where beneficial mutations are deleterious mutations occur at genome-wide rates Ub and Ud, rare, and occur one-by-one in discrete selective sweeps (Leigh respectively. We define U ¼ Ub þ Ud as the total mutation 1970, 1973; Painter 1975; Gillespie 1981; Ishii et al. 1989; rate, and e [ Ub=Ud as the ratio of beneficial to deleterious Kessler and Levine 1998; Johnson 1999b; Tanaka et al. mutation rates. We focus on adapting populations, and con- 2003; Andre and Godelle 2006; Wylie et al. 2009; Desai and sider both the successional-mutations regime (where benefi- Fisher 2011). However, our understanding is much more lim- cial mutations are rare), and the clonal interference regime. ited in larger populations, where beneficial mutations are For most of our analysis, we assume that beneficial mutations more common, and multiple adaptive lineages must compete all confer the same fitness advantage sb; though we also comment with each other for fixation. This is the regime where the in- on how our results can be extended to the case where beneficial direct effects of linked selection are likely to be maximized, mutations have a distribution of fitness effects. We also assume and it is also the most relevant for understanding the microbial that there is no macroscopic epistasis among beneficial mutations evolution experiments where mutation rate modifiers have [as defined by Good and Desai (2015)], so that Ub and sb remain been observed to spread. In this clonal interference regime, fixed as the population adapts. On sufficiently long timescales, even the most basic questions about this process remain unan- this assumption may eventually break down, so we also comment swered: exactly how large can the deleterious load be before it on the effects of changing Ub or sb in the Discussion. effectively selects against a mutator allele? And how does this We make no specific assumptions about the fitness effects depend on the population size, the supply of beneficial muta- of deleterious mutations, other than that they are purgeable, tions, and the magnitude of the mutator phenotype? i.e., deleterious mutations are sufficiently costly that they are Here, we address these questions in the context of a simple unlikely to fix, and do not significantly reduce the overall rate model of adaptation, which is motivated by recent microbial of adaptation of the population. We describe the technical evolution experiments. In particular, we focus on a regime in conditions required for this to be true, and the qualitative which deleterious mutations have a negligible impact on the implications of nonpurgeable deleterious mutations in more rate of adaptation, even though they may have a large effect on detail in the Discussion. the fates of mutation rate modifiers. Within this model, we Our goal is to analyze the fate of a modifier allele that calculate the fixation probability of an allele that changes the changes the overall mutation rate U by a factor r (while

1250 B. H. Good and M. M. Desai leaving e and sb unchanged). For simplicity, we assume that pest ¼ 2sb: SelectivesweepsthenoccurasaPoissonpro- this allele has no direct influence on fitness, although in the cess with rate Discussion we show how this assumption can be relaxed. =t : These modifier alleles are produced at rate m from the 1 est ¼ 2NUbsb (1) wild-type background, and we assume that m is sufficiently When a successful beneficial mutation establishes, it small that the modifiers can be treated independently. In this starts to grow deterministically and requires roughly case, the substitution rate of the modifier is Nmp ðrÞ; where fix T ¼½ð2=s ÞlogðNs Þ generations to complete its sweep. p ðrÞ is the fixation probability of a single modifier individ- fix b b fix By definition, this time is small compared to the time between ual. A natural measure of how selection “favors” or “disfa- sweeps, so the rate of adaptation is just y ¼ s =t : A modifier vors” the modifier can be obtained from the scaled fixation b est allele would increase or decrease this rate by factor of r,ifit probability, Np ðrÞ: When Np . 1; the allele is favored by fix fix was able to fix. We now analyze the dynamics of this process. selection, since it substitutes more rapidly than a neutral When a modifier allele first arises, it starts at an initial allele; conversely, the allele is disfavored when Np , 1: This fix frequency f ð0Þ¼1=N; and it will then drift neutrally until scaled fixation probability will be our primary focus in the m the next selective sweep occurs. Most of the time, the lineage analysis below. In particular, we are interested in determining will fluctuate to extinction within the first few generations. the parameters where Np transitions between favored fix However, with probability 1=t; it will survive for t gener- (Np . 1) and unfavored (Np , 1). fix fix ations, and reach size f ðtÞt=N (Fisher 2007). In the ab- We are primarily interested in situations applicable to m sence of selective sweeps, the modifier lineage can only fixby microbial populations, and particularly to laboratory drifting to fixation; this happens with probability 1=N, and evolution experiments. This motivates certain technical occurs over a timescale of order N generations. Thus, assumptions we describe in the analysis below. It also depending on how this drift timescale N compares with the motivates our focus on asexual populations where recombi- sweep timescales t and t =r; two distinct types of fixation nation can be neglected. This asexual case is particularly est est dynamics can emerge. relevant because, when recombination is absent, modifiers When t and t =r are both much larger than N, then the of mutation rate remain perfectly linked to the beneficial or est est fate of the modifier lineage will be determined long before deleterious mutations they generate, and hence experience the next selective sweep occurs. Drift is the dominant evolu- the strongest indirect selection pressures. In principle, we tionary force, and p ðrÞ¼1=N: could also apply our analysis to physically linked regions fix In contrast, if either t or t =r are small compared to N, within sexual (“linkage blocks”) that are unlikely est est the fixation process is instead controlled by genetic draft. In to recombine on the appropriate timescales (Neher et al. this case, the modifier lineage will fix only if it is lucky enough 2013; Good et al. 2014; Weissman and Hallatschek 2014). to produce and hitchhike with the next selective sweep. Since However, the scale of these linkage blocks is a complex we are primarily interested in understanding these effects, problem, and understanding these effects is beyond the we will focus on this regime for the remainder of our analysis. scope of this study. The next selective sweep is produced by two competing In the remaining sections, we calculate Np ðrÞ as a func- fix Poisson processes, corresponding to the beneficial mutants tion of the underlying model parameters, and then we use from the wild type and modifier lineages. These respectively these predictions to analyze the dynamics of mutation rate produce sweeps at rates evolution.

Data availability l0ðtÞ ¼ 2NUbsb½1 2 fmðtÞ; (2a) The authors state that all data necessary for confirming the lmðtÞ¼2NU s r fmðtÞ: (2b) conclusions presented in the article are represented fully b b within the article. The probability that the next sweep is produced by the modifier lineage is then simply Z R The Successional Mutations Regime N t 2 l0 t9 þlm t9 dt9 l 0 ½ ð Þ ð Þ ; ; pfix ¼ mðtÞe dt (3) To gain intuition into the forces that determine NpfixðrÞ it 0 is useful to start by considering the simplest case, where deleterious mutations can be neglected, and adaptation where the angled brackets denote an average is over the proceeds by a sequence of discrete selective sweeps. This random lineage trajectory fmðtÞ: In the regime we are consid- “successional mutations” regime applies whenever bene- ering, the modifier lineage will remain rare until the next ficial mutations are sufficiently rare that they seldom seg- selective sweep occurs. The trajectory fmðtÞ is therefore de- regate within the population at the same time. This scribed by the Langevin dynamics, will be true provided that NUblogðNsbÞ1(Desaiand rffiffiffiffiffi fi @ Fisher 2007). In this regime, bene cial mutations are fmðtÞ fmh ; @ ¼ ðtÞ (4) produced at rate NUb and establish with probability t N

Mutators in Rapidly Adapting Populations 1251 where hðtÞ is a Brownian noise term (Gardiner 1985). The When a new beneficial mutation occurs, the wild type will be in solution of Equation (3) can be complicated, since both the mutation-selection balance with respect to its deleterious muta- overall rate of sweeps and their probability of arising in tions, so the mean fitness of the population is 2Ud: Even if it arises the modifier lineage depend on the random trajectory fmðtÞ: on a mutation-free background, the new beneficial lineage will We present an exact solution of these equations in Appendix continue to produce loaded individuals at rate Ud: This exactly A, using techniques developed by Weissman et al. (2009). cancels the growth advantage of the mean fitness of the popu- For the present purposes, however, it will be more useful to lation, so that the establishment probability is still just focus on an approximation. If the modifier lineage stays pest ¼ 2ðsb þ Ud 2 UdÞ¼2sb: Since most genetic backgrounds sufficiently rare that it does not influence the overall rate of are in this unloaded state, the overall rate of sweeps is unchanged fi fi sweeps, then we can approximate l0ðtÞþlmðtÞ1=test: from before. This means that if a modi er allele were to x, it Equation (3) then reduces to would still change the rate of adaptation by a simple factor of r. However, while the deleterious mutations do not affect the Z N 2 =t t est rate of adaptation, they can still have a large effect on the e : pfix ¼ rh fmðtÞi dt (5) modifier lineage while it is rare. The modifier produces 0 test doomed lineages at rate rUd; which is not completely can- In other words, the fixation probability is given by the average celled by the mean fitness of the wild type. Instead, mutator size of the lineage at the time of the next sweep, and this time is alleles will feel an effective cost of magnitude Udðr 2 1Þ; unaffected by the presence of the modifier. In this case, the while antimutators will experience an effective fitness advan- modifier lineage is neutral and h fmðtÞi ¼ 1=N; so that tage. Generalizing Equation (4), the mutation-free portion of the modifier lineage will then evolve as : NpfixðrÞ¼r (6) rffiffiffiffiffiffiffiffiffiffi @fmðtÞ fmðtÞ Thus, we see that the modifier fixation probability (like the ¼ 2 U ðr 2 1ÞfmðtÞþ hðtÞ: (7) @t d N proportional change in the rate of adaptation) is independent of the population size, the mutation rate, and the fitness Beneficial mutations produced by this lineage will likewise benefits of the driver mutations. have an effective fitness sb 2 Udðr 2 1Þ; and will establish with To derive Equation (6), we made the approximation that probability l0ðtÞþlmðtÞ1=test: Substituting our expressions for l0ðtÞ pm ¼ 2ðs 2 U ðr 2 1ÞÞ: (8) and lmðtÞ in Equation (2), we see that this is equivalent to the est b d assumption that ðr 2 1ÞfmðtÞ1: However, since fm and t are both random variables, we cannot determine the validity of In exactly the same manner as before, the next selective sweep this assumption based on the average values alone. We must is produced by two competing Poisson processes with rates instead consider the typical dynamics that contribute to the =t ; 1 average in Equation (5): with probability 1 est the modi- l 2 ; 0ðtÞ¼t ½1 fmðtÞ (9a) fier lineage survives for test generations and reaches size est fmðtestÞtest=N: This ensures that the average size of the   t = =t = ; r U ðr 2 1Þ lineage is just ½ð est NÞð1 estÞ ð1 NÞ as expected. It l 2 d ; mðtÞ¼t 1 fmðtÞ (9b) also shows that the typical values of ðr 2 1ÞfmðtÞ will remain est sb small compared to one provided that rt N: For simplic- est fi ity, we will assume that this condition holds for the remainder and the xation probability is given by the average in Equation (3). fi of our analysis. For moderate r, this is actually the same as- If we again make the assumption that the modi er lineage has a negligible influence on the overall sweep rate sumption we already made in focusing on the genetic-draft (lm þ l0 1=test), this reduces to regime. For large r, it imposes an upper limit r ðN=testÞ1 on the range of modifier effects that we consider. However,   Z N 2 =t U ðr 2 1Þ e t est this is not a fundamental problem for the analysis; we con- p ¼ r 1 2 d h f ðtÞi dt: (10) fix s m t sider such large-effect modifiers in Appendix A. b 0 est

Incorporating purgeable deleterious mutations In this case, the average lineage size is now h fmðtÞi ¼ 2 2 e Udðr 1Þt=N; and we obtain A similar picture applies in the presence of strongly deleterious   mutations. In particular,we assume that the typical fitness costs 2 2 Udðr 1Þ are larger than s ; so that a driver mutation can fixonlyina r 1 b sb : background that is free of deleterious mutations. If the costs NpfixðrÞ¼ (11) 1 þ Udtestðr 2 1Þ are also much larger than Ud; then the vast majority of the population will be of this mutation-free type. We refer to these The validity of this expression depends on our assumption that as purgeable mutations, and we will assume that all deleteri- l0ðtÞþlmðtÞ1=test: In the same manner as above, we can ous mutations are purgeable in the analysis that follows. establish the region of validity of this approximation by

1252 B. H. Good and M. M. Desai considering the typical dynamics behind the averages in Equation (10). These will depend on the sign of Udðr 2 1Þ: For a mutator allele (r . 1), the modifier lineage will feel an effective fitness cost Udðr 2 1Þ: If this cost is small com- pared to 1=test; then selection will barely have time to influ- ence fmðtÞ before the next sweep occurs, and mutators will continue to fix with probability Npfix r: On the other hand, if Udðr 2 1Þ1=test; then this fitness cost exponen- tially suppresses the probability that the mutator survives until the next sweep. In particular, the mutator will survive for t 1=Udðr 2 1Þ generations with probability 2Udðr21Þt Udðr 2 1Þe ; and, provided that it does so, it will no longer grow indefinitely, but will instead reach a maxi- mum size of order 1=NUdðr 2 1Þ (Desai and Fisher 2007). If the time between sweeps was deterministic, this would sim- Figure 1 The fixation probability of a mutation rate modifier in the suc- 2 2 t cessive mutations regime. Solid lines depict Equation (11) for four sets of ply suppress the fixation probability by a factor e Udðr 1Þ est : parameters, which illustrate the four characteristic shapes of NpfixðrÞ: In However, this is not actually the case, since the next sweep is 7 22 24 all four cases, the base parameters are N ¼ 10 ; sb ¼ 10 ; U ¼ 10 , generated by the random Poisson process described above. In and e ¼ 1025; with modifications listed in the inset. particular, the random nature of this process will occasionally produce a sweep that occurs much earlier than test: Such anom- alously early sweeps are uncommon, occurring with probability of Udtest; Equation (11) takes on one of two characteristic t=test 1; but they are far more common than a mutator shapes. If Udtest . 1; the fixation probability is a monotoni- lineage that survives for a typical sweep time. Combining these cally decreasing function of r (see Figure 1), and mutators facts, we find that the mutator fixation probability is domi- will never be favored to invade. Instead, antimutators will be nated by anomalously early sweep times for which positively selected. Note that this happens even though the t 1=Udðr 2 1Þtest ; which leads to a power law decay antimutator actually causes a decrease in the overall rate of NpfixðrÞ}1=Udtestðr 2 1Þ: For a strong-effect mutator (r 1), adaptation. this power-law decay exactly balances the lineage’sincreased On the other hand, for Udtest , 1; the fixation probability beneficial mutation rate, so that the r-dependence of pfix is medi- takes on a paraboloid shape (Figure 1). It crosses NpfixðrÞ¼1 m : at exactly two points: once at r 1 and once at ated primarily through the reduced establishment probability, pest ¼ In contrast, antimutators will have an effective fitness advan- sb 2 t : tage relative to the wild type, of magnitude Udð1 2 rÞ: If r* ¼ ð1 Ud estÞ (13) Ud Udð1 2 rÞ1=test; this fitness advantage will still have a negli- fl : 2 t ; gible in uence on NpfixðrÞ However, when Udð1 rÞ est 1 When r* . 1; modifiers will be favored in the range 2 the antimutator will establish with probability Udð1 rÞ,andwill 1 , r , r*; and disfavored elsewhere (i.e., some mutators will Udðr21Þt start to grow deterministically as fmðtÞe =NUdðr 2 1Þ: be favored). Conversely, when r* , 1; modifiers will be fa- Again, if the time to the next sweep was deterministic, this would vored in the range r* , r , 1(i.e., some antimutators will be result in a simple exponential enhancement of Npfix: But because favored). It is also useful to rewrite r* in terms of the muta- t is exponentially distributed, it interacts with the exponential tion rate Um ¼ Ur of the modifier lineage: growth of the lineage in a strong way. In particular, as  1 Udð1 2 rÞtest approaches 1, the average in Equation (11) will * [ e 2 : Um Ur* ¼ sbðÞ1 þ 1 e (14) be dominated by anomalously late sweep times for which fmðtÞ 2 N sb is no longer rare, and our approximation that l0 þ lm 1=test Note that this is always positive (since 1=2Ns e , 1), and it is breaks down. Thus, Equation (11) will only be valid provided b independent of U. that 1 2 Udð1 2 rÞtest 1=logðN=testÞ: Outside this regime, the antimutator lineage will appear to sweep to fixation on its own, Dynamics of mutation rate evolution without the help of an additional driver mutation. We are now in a position to analyze how mutation rates evolve The fixation probability in Equation (11) is illustrated in over time. Imagine that we start with some particular values of Figure 1 for several choices of parameters. Its overall shape is N, U, e, and sb: This will correspond to a particular value of determined by the key parameter * test and Um: If Udtest . 1; mutators are disfavored and anti- mutator alleles will be positively selected instead. After an t 1 ; Ud est ¼ (12) antimutator allele fixes, both U and U are reduced, but 2eNsb b d Udtest is unchanged, so natural selection will continue to which depends only on N, sb; and e [ Ub=Ud; and is indepen- select for lower mutation rates until this force is balanced dent of the overall mutation rate U. Depending on the value by drift, mutational pressure, or other physiological costs.

Mutators in Rapidly Adapting Populations 1253 * On the other hand, if Udtest , 1 and U , Um; mutators will opposite can actually be true: clonal interference enhances be favored provided that their resulting mutation rate is less the fixation of mutator alleles, even as they provide a dimin- * than Um: If a mutator allele fixes, Ub and Ud will be increased, ishing overall benefit for the rate of adaptation. * but Udtest and Um will remain constant. Thus, natural selec- In the following subsection, we introduce a traveling-wave tion will continue to favor increased mutation rates until U formalism for calculating NpfixðrÞ in the presence of clonal reaches a special value, interference. Before doing so, however, it will be useful to  consider this process from a heuristic perspective. This will ^ * e 2 1 ; allow us to identify the key forces and parameters involved, U ¼ Um ¼ sbðÞ1 þ 1 e (15) 2 Nsb and will provide intuition for the more rigorous analysis that follows. which is stable against further changes in the mutation rate. If instead the initial mutation rate U . U^; antimutators will be Heuristic analysis . ^: favored provided that their resulting mutation rate is U In the absence of deleterious mutations, clonal interference Natural selection will then continue to favor decreased mu- alters the dynamics of fixation in two mainways. First, success- tation rates until U reaches U^: We describe these dynamics in ful mutations can only occur in the most highly fit genetic more detail in the Discussion. backgrounds in the population. These individuals lie in the Of course, this analysis crucially depends on the assump- extreme right tail, or “nose,” of the population fitness distri- tion that any mutation rate increases will still lie within the bution, which steadily increases in fitness as the population successive mutations regime. This will be true provided that adapts. In the regime described by Equation (17), the relative NU^ eNs 2 1 1=logðNs Þ: But, in order for a nonzero b b b fitness of the nose is given by stable mutation rate to exist in the first place, we previously   required that eNsb 1=Udtest . 1: These two conditions can 1 sb be satisfied only if xc log ; (18) test Ub   1 1 1 , e , 1 þ : (16) which is much larger than the size of a single driver mutation. 2Nsb 2Nsb logðNsbÞ These individuals have already acquired q ¼ xc=sb 1 more adaptive mutations than the average individual, but they are This is an extremely small region of parameter space, and it not yet destined to fix. The reason is that there are still grows increasingly narrow as Ns increases. As a result, the b enough individuals in the nose that they will collectively pro- stable mutation rate in Equation (15) is actually a rapidly duce multiple additional driver mutations in the next t varying function of N, s ; and e. For larger values of e that est b generations, and these will occur in a relatively narrow time violate the stringent constraints in Equation (16), mutator window Dt t (Desai and Fisher 2007). By definition, all alleles will generically drive mutation rates into regimes est but one of these mutations must eventually be outcompeted. where selective sweeps begin to interfere with each other. But this means that the process of fixation within the nose We now turn to an analysis of this case. takes place over multiple establishment intervals, each of length test: The Clonal Interference Regime Since xctest 1; genetic drift is not directly relevant for fi fl When multiple beneficial mutations segregate at the same most of this xation process. Random uctuations in the line- time, many potential drivers are lost due to competition with age sizes are still important, but these are now driven by other, fitter genetic backgrounds. This reduces the rate of genetic draft, which arises from slight differences in the rela- successful drivers in a way that depends on the relative values tive order of the next round of driver mutations. During most of this process, the lineages founded by different driver mu- of Nsb and NUb: In the regime most relevant for our current tations make up less than 1=q of the total size of the nose. study, Desai and Fisher (2007) have shown that test is re- duced to However, approximately once every q establishments, an anomalously early driver mutation will occur and reach an 1 2s logðNs Þ Oð1Þ fraction of the nose (Desai et al. 2013). This roughly b b ; (17) t 2 coincides with a fixation event. In order to fix, a lineage must est log ðsb=UbÞ therefore (i) arise in the nose, (ii) persist for q additional which is valid provided that 1 logðsb=UbÞ2logðNsbÞ, establishment intervals, and (iii) be lucky enough to hitch- 2 “ ” and 2logðNsbÞlog ðsb=UbÞ: Similar expressions can be hike with the special jackpot driver event. obtained for other parameter regimes, all of which share Mutation-rate modifiers can be incorporated into this pic- t the weak dependence on NUb (Fisher 2013). Since adap- ture in a straightforward way.The key quantities est and q are tation is no longer mutation-limited, one might guess that only weakly dependent on the mutation rate. Provided that mutators will be less strongly favored in this regime. How- jlogðrÞj logðsb=UbÞ; the modifier versions of these param- ever, previous simulation studies (Tenaillon et al. 1999) and eters will be similar to the wild type, and the competition heuristic reasoning (Desai and Fisher 2011) suggest that the between these lineages will play out according to the same

1254 B. H. Good and M. M. Desai dynamics as above. The modifier is no more or less likely to arise in the nose compared to a neutral mutation. However, provided that it does occur in one of these special genetic backgrounds, a mutator lineage is r times more likely than a neutral lineage to generate an additional driver, while an antimutator lineage is r times less likely to do so. Since the modifier lineage must generate q additional drivers to fix, its overall fixation probability is given by

q NpfixðrÞr : (19)

Thus, we see that clonal interference increases the fixation prob- ability of mutators by q factors of r. This increase can be sub- stantial for large r,evenwhenq 2(seeFigure2).Fromour discussion above, we see that this increase is primarily driven by the fact that multiple additional drivers are required for fixation. In the presence of deleterious mutations, a mutator lineage again feels an effective cost of Udðr 2 1Þ; so it will tend to decline in frequency relative to the other individuals in the nose. In particular, when the next burst of driver mutations arises, the mutator lineage will have decreased in frequency 2 2 t by a factor of e Udðr 1Þ est ; which makes it that much less likely to survive to the next round. Similarly,an antimutator lineage will have increased in frequency by an analogous factor of 2 t eUdð1 rÞ est : The overall fixation probability then becomes fi fi   Figure 2 (Top) The xation probability of a mutation rate modi er in q the clonal interference regime when Ud ¼ 0: Symbols denote the results 2ðr21ÞUdtest Np ðrÞ re : (20) 22 fix of forward-time solutions (described in Appendix C) for sb ¼ 10 ; 25 5 7 9 Ub ¼ 10 ; and N 2 f10 ; 10 ; 10 g Solid lines denote the theoretical We discuss the regimes of validity of this expression in our predictions in Equation (38). For comparison, the successive muta- Np r Np more rigorous analysis below.For the purposes of our heuristic tions prediction ( fix ) and neutrality ( fix 1) are shown by the dashed lines. (Bottom) The rate of adaptation of a successful modifier analysis, it will be more useful to focus on the implications of lineage, relative to that of the wild type, for the same set of popula- Equation (20). tions above. The solid black line denotes the asymptotic prediction, 2 2 Similar to the selective sweeps case, the direction of selec- vðrÞ=v log ðsb=UbÞ=log ðsb=rUbÞ; which is independent of N. For com- tion in Equation (20) is again determined by the product parison, the successive mutations prediction (vðrÞ=v r), and no change (vðrÞ=v 1), are shown by the dashed lines. Udtest: There are three characteristic regimes of behavior. When Udtest 1; mutators will be favored provided that 1 þ e Um ¼ Ur is less than a maximum value, ^ : U ¼ t (24)   est 1 þ e 1 U* ¼ log ; (21) ^; t m t t Note that this is actually an implicit relation for U since est est Ud est ^ ^ depends on Ub ¼ eU=ð1 þ eÞ: Substituting our expression for ^ which corresponds to the largest value of r for which test in Equation (17), and solving for U; we find that Np ðrÞ . 1: Conversely, when U t 1; antimutators will fix d est 2s logðNs Þ be favored above a minimum value U^ b ÀÁb ; (25) 2 1  log e 1 þ e 2 t * t Ud est : Um ¼ Ud este (22) where the region of validity for Equation (17) [and hence test Equation (25)] now becomes

Similar behavior is obtained for Udtest is close to one, except pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi * 1 2log Ns log 1=e 2log Ns : (26) that Um is now given by ð bÞ ðÞ ð bÞ   1 2 U t This is still a restrictive parameter range for e, although it is U* ¼ U 1 þ d est : (23) m 2 much broader than Equation (16) in the successive mutations regime, and it grows larger with increasing Nsb: When Udtest ¼ 1; the range of favorable modifiers vanishes, Provided that these conditions are met, we see that the and mutators and antimutators are both selected against. The stable mutation rate in Equation (25) is only weakly depen- evolutionarily stable mutation rate is therefore given by dent on N and e, and is much more strongly influenced by sb:

Mutators in Rapidly Adapting Populations 1255 This contrasts with the behavior we found in the successive solution (based on ideas introduced in earlier studies by mutations regime. In addition, we see that, unlike in the Tsimring et al. (1996), Neher et al. (2010), and Hallatschek successive mutations regime, the stable mutation rate (U^) (2011)), which is thought to be relevant for many microbial does not necessarily coincide with the maximally permitted evolution experiments (Desai et al. 2007; Perfeito et al. 2007; * ^ mutator or antimutator allele (Um) when U 6¼ U: This can Wiser et al. 2013; Barroso-Batista et al. 2014). In this regime, have important consequences for the dynamics of mutation the fitnesses of unloaded individuals are approximately nor- rate evolution in this regime. If the mutation rate starts far mally distributed, ^; fi above or below U natural selection can favor modi er alleles À Á   1 2x2 that overshoot the stable value by a substantial amount, f x~ pffiffiffiffiffiffiffiffiffi e 2y u xc 2 x~ ; (29) which can lead to a nonmonotinic approach to the stable 2py point. We will revisit these dynamics in more detail in the with a sharp cutoff at x~ ¼ x : This corresponds to the “nose” of Discussion. c the fitness distribution described above. Meanwhile, the sur- Formal analysis vival probability wðx~Þ can be approximated by the piecewise form We now turn to a more formal derivation of the results de- scribed above. To do so, we make use of “traveling-wave” 8 > 2~x if x~ . xc; formalism developed in previous work (Tsimring et al. <> À Á 22 2 ~ x xc 1996; Rouzine et al. 2003; Neher et al. 2010; Hallatschek w x 2x e 2y if x 2 s # x~ # x ; (30) > c c b c 2011; Neher and Shraiman 2011; Good et al. 2012; Fisher : 0if x~ # x 2 s : 2013; Good and Desai 2014). This formalism focuses on the c b fi distribution of relative tness within the population, which In this context, x can also be thought of as an interference ; fi c we denote by fðxÞ and the xation probability of a (wild- threshold. Lineages that are at relative fitness x . x will fix fi c type) individual with relative tness x, which we denote by provided they survive drift and establish; this occurs with wðxÞ: In the absence of mutator or antimutator alleles, we probability 2x: Below xc, the fixation probability drops off and others have previously shown that fðxÞ and wðxÞ satisfy rapidly because lineages can establish but still be lost to in- the partial differential equations, terference. Once a lineage is more than sb below xc; it can essentially never catch up with the remaining lineages at the @f 2y ¼ðx 2 U 2 U Þf þ U fðx 2 s Þ; (27a) nose, and it will therefore have a negligible fixation proba- @x d b b b bility. The location of xc is determined by the auxiliary @ 2 condition y w 2 2 2 w ; ¼ðx Ud UbÞw þ Ubwðx þ sbÞ (27b)  2 @x 2 1 s2 Ub sb xcsb2 b 1 12 e y 2y : (31) s x where y [ sb=test is the average rate of adaptation (Good and b c Desai 2014). Note that Equation (27) implicitly assumes that After substituting our expressions for fðxÞ and wðxÞ into the deleterious mutations are purgeable, i.e., they do not fix, Equation (28), the self-consistency condition becomes and constitute a negligible fraction of the population. The rate of adaptation (or equivalently, t [ s =y) is set by the x2 est b 2 c 2x s e 2y 1 self-consistency condition, pc ffiffiffiffiffiffiffiffiffib : (32) Z 2py N 1 fðxÞwðxÞdx ¼ ; (28) N Together with Equation (31), this completely determines v and xc as a function of the underlying parameters. Asymptotic which is just another way of saying that the fixation prob- formulae for these quantities are given in Equations (17) and ability of a neutral mutation must be equal to 1=N: For a (18); more accurate estimates can be obtained by solving detailed derivation of these equations, see Good and Desai Equations (31) and (32) numerically. (2014).Notethat,becausewehaveassumedthatdeleteri- We have previously shown that this solution applies when- pffiffiffi pffiffiffi ous mutations are purgeable, Ud enters into Equations (27) ever xc 2 sb y; sb y; and Ub sb (Good and Desai and(28)onlyasanoverallshiftinthemeanfitness of the 2014). The first of these conditions says that the parents of population. If we measure fitnesses using the shifted vari- successful lineages must be highly fit, i.e., that clonal inter- ~ able x ¼ x 2 Ud (i.e., fitness relative to the average unloaded ference is indeed present. The second condition, which also individual), then the evolution equations revert back to the implies the third, says that the vast majority of individuals in purely beneficial case. We will adopt this convention from the population have roughly the same number of driver mu- now on. tations, i.e., mutation pressure is not too strong. In terms Even in the simplified model of Equation (27), there are of the underlying parameters, these conditions become 2 many possible parameter regimes that one may consider 1 logðsb=UbÞlogðNsbÞlog ðsb=UbÞ; as we stated after (Fisher 2013). Here we will focus on a particular approximate Equation (17) above.

1256 B. H. Good and M. M. Desai Once we have set up this traveling-wave formalism, This places an upper limit on the range of modifier effects mutation-rate modifiers can be added in a straightforward that we can consider. But since sb=Ub 1; this includes many way.Since the fate of any allele is determined whileit is rare, we realistically large modifiers of order r 100: can neglect the effects of the modifier on the wild-type pop- Given our solution for wmðx~Þ; we can compute the mar- ulation, so that fðxÞ and v are unchanged from above. When a ginal fixation probability of the modifier by averaging over modifier allele occurs, it will arise on a genetic background all the fitness backgrounds that the modifier could have whose fitness is drawn from fðxÞ: We then introduce a new arisen on: function, wmðxÞ, describing the fixation probability of the mod- Z À Á À Á ifier allele as a function of its initial fitness. This function satisfies ~ ~ ~ pfixðrÞ ¼ f x wm x dx: (37) a similar equation as wðxÞ; except with a different mutation rate:

@ h i We evaluate this integral in Appendix B. We find that y wm ~ 2 2 2 @~ ¼ x ðr 1ÞUd rUb wm x n 2 (33) q½logðrÞ2Udtestðr21Þ 2 ½logðrÞ2Udtestðr21Þ Np r e 3 e 2 À Á w2 fixð Þ þ rU w x~ þ s 2 m: t 2 b m b eUd estðr 1Þ 2 1 2 3 ; (38) Udtestðr 2 1Þ Thus, in addition to the increase in Ub; we see that a mutator fi 2 lineage experiences an effective tness cost Udðr 1Þ corre- where we have employed the short-hand notation q ¼ xc=sb; sponding to its increased deleterious load. This cost enters as a test ¼ sb=y; and n ¼ 1=sbtest: In the asymptotic regime we are ~ shift in the overall location of wmðxÞ; which we can account for considering, 1=q and n are both small parameters. To leading ~ ~ ~ by defining a shifted function, wmðxÞ¼wmðx 2 Udðr 2 1ÞÞ: order, Equation (38) converges to the simpler expression After this transformation, Equation (33) is of the same form from our heuristic analysis, which we now recognize as the as Equation (27b). Then, provided that the modifier mutation technical statement that rate rUb is still in the same regime as Ub; the solution for the ~ ~ ~ shifted function wmðxÞ will have the same form as wðxÞ; logNpfixðrÞ lim ¼ logðrÞ 2 Udtestðr 2 1Þ: (39) 8 q/N q > 2~x if x~ . x ; n/0 <> cm À Á 22 2 ~ ~ x xcm w x 2y 2 # ~ # ; (34) m ¼ > 2xcme if xcm sb x xcm :> However, since the errors are multiplied by a large number q ~ 0if x # xcm 2 sb; and exponentiated, the full version in Equation (38) is often required for quantitative accuracy. ; except with a different interference threshold, xcm which We illustrate the full expression in Equation (38), and satisfies compare it to forward-time simulations in Figure 2 and Fig-

 2 ure 3. The accuracy is generally good, although there are 1 s2 rUb sb xcmsb2 b some systematic deviations for large r when the effects of 1 ¼ 12 e y 2y : (35) sb xcm the deleterious load are particularly costly. These are ulti- mately the result of two factors: (i) inaccuracies in our ap- ; Like the wild-type interference threshold xc the location of xcm is proximate solution for wðxÞ for x 2 s , x , x , and (ii) the : ; c b c independent of Ud Since mutators are favored when Ud ¼ 0 we fact that logðrÞ is starting to approach logðs =U Þ: More accu- , . : b b expect that xcm xc whenever r 1 In other words, we expect rate approximations for wðxÞ derived by Fisher (2013) could that mutators have a lower interference threshold, since these be used to improve the quantitative accuracy for these pa- fi lineages generate bene cial mutations more rapidly and, once in rameters; we leave this for future work. the nose, are therefore less likely to be lost to clonal interference. In order for this solution to apply, it must satisfy the same pffiffiffi conditions as the wild-type population: xcm 2 sb y and Discussion rU s : This will be true provided that x is still close b b cm In rapidly adapting populations, the fates of mutator and to x ; or, equivalently, that the fractional difference c antimutator alleles depend on a careful balance between d ¼ x =x 2 1 is small compared to 1. Given this assumption, x cm c the benefits of hitchhiking and the cost of a deleterious load. we can divide Equation (35) by Equation (31), and solve for d : x Our analysis shows how the fixation probabilities of these 2 3 mutation-rate modifiers depend on the population size, the y 6 d 7 y beneficial and deleterious mutation rates, and the strength of d 2 6 rð1 þ xÞ 7 2 : x ¼ log4 d 5 logðrÞ (36) selection. By searching for parameters where mutators and xcsb xc x xcsb 1 þ antimutators are both disfavored by selection, we can identify xc 2 sb particular mutation rates, U^, that are stable under future Since xcsb=y logðsb=UbÞ in this regime, we see that this ap- evolution. At these mutation rates, the benefits of hitchhiking proximation will hold provided that jlogðrÞj logðsb=UbÞ: are exactly balanced by the costs of the deleterious load, so

Mutators in Rapidly Adapting Populations 1257 that neither mutators nor antimutators are favored. We now consider how a population will approach the stable mutation rate, and the implications of this mutation rate evolution for the rate of adaptation within the population. Finally, we de- scribe the approximations we have made throughout our analysis and the resulting limitations of our approach. Approach to the stable mutation rate Consider a situation in which the population starts with a ^ mutation rate U0 6¼ U; and fixes a sequence of modifier alleles with mutation rates U1; U2; ...; Un: What are the typical val- ues of this mutation rate trajectory, under the hypothesis that each step was favored by selection? In both the successional mutations and clonal interference ^ regimes, we found that U depends only on N, sb; and e. This implies that the mutation rate will always evolve toward this Figure 3 The fixation probability of a mutation rate modifier in the ^ U . : unique stable point: Un/U: However, the manner in which clonal interference regime when d 0 Symbols denote the results 7 the population approaches U^ can be strongly influenced by of forward-time simulations with the base parameters N ¼ 10 ; 22 25 24 21 sb ¼ 10 ; Ub ¼ 10 ; Ud ¼ 10 ; and sd ¼ 10 ; with modifications clonal interference. listed in the figure. Solid lines denote the theoretical predictions in In the successional mutations regime, selection favors a Equation (38). monotonic approach to U^: Consider, e.g., a mutation rate ^ U0 , U: In this case, we have shown that mutators will be ^ able to fix, provided that the new mutation rate Um ¼ rU lies If the population starts at a mutation rate U0 U; then ^ in the range U0 , Um , U: When one of these alleles fixes, the only modest changes in the mutation rate will be favorable. ^; new mutation rate will be given by U1 ¼ Um; and the process However, if the population starts at a mutation rate U0 U will repeat itself. Mutators will still be favored to fix provided Equation (40) shows that mutators will be positively se- ^ ^ ^ that U , U , U; which will lead to a U . U ; and so on. lected, provided that U0 , Um , UlogðU=U0Þ: While the most 1 m 2 1 ^ Thus, evolution will favor a monotonic sequence of mutator strongly beneficial modifier has Um ¼ U; the range of favored ^ ^ alleles until Un ¼ U; after which the mutation rate will be mutators also includes values of Um that are larger than U by ^= : stable to further changes. An exactly analogous conclusion a factor logðU U0Þ1 This means that selection can fa- ^ holds if U0 starts above U : here, selection will tend to fix vor mutator alleles that overshoot the stable mutation ^ antimutator alleles that lie in the range U , Um , Ui until rate by a substantial amount. If a mutator allele of this ^ maximal strength fixes, the new mutation rate will be Un ¼ U: In both cases, mutator or antimutator alleles that ^ ^= . ^; “overshoot” the stable mutation rate are always disfavored. U1 ¼ UlogðU U0Þ U and selection will immediately favor the fixation of antimutator alleles, despite the fact that the This is true even if such alleles would move the mutation rate ^ closer to the stable rate; e.g., an allele that takes a population location of U has not changed (see Figure 4 and Figure 5). ^; from 0:1U^/1:1U^ would not be positively selected. If U U these antimutator alleles can overshoot the When clonal interference is present, the approach to the stable mutation rate in the other direction, by a factor =^ 2U=U^ : stable mutation rate is more complex. This is easiest to see if U Ue 1 In the example above, this could then lead ^= ; ; ^: to U2 ¼ U0logðU U0Þ which is larger than the initial muta- we rewrite Equation (20) in terms of Um U, and U Provided ^ ^ tion rate U0 but still much smaller than U: Thus, while the that jlogðUi=UÞj logð1=eÞ; the quantities testðUÞ and qðUÞ will stay roughly constant over the relevant range of muta- population will still approach the stable mutation rate over ^ ^ time, this approach does not have to be monotonic, even in tion rates, and we can approximate testðUÞtestðUÞ [ 1=U: the absence of epistasis or environmental variation. More- This yields a simple heuristic formula, over, the functional form of Equation (40) implies that this ! q 2U =U^ behavior is highly asymmetric: an antimutator allele can Ume m ^ ; overshoot U by a larger amount (and with a larger Npfix) than NpfixðUmÞ 2 =^ (40) Ue U U a mutator allele with the same value of jlogðU=U^Þj: This can ^ ^ also be seen in Figure 4 and Figure 5. which depends only on the scaled mutation rates U=U; Um=U; and q. A more accurate (but more cumbersome) version can Application to a long-term evolution experiment in E. coli be obtained by substituting the full expressions for testðUÞ and qðUÞ into Equation (38). We illustrate this full expression Several studies have observed the spread of mutator alleles in in Figure 4 for a particular combination of Nsb and e, although microbial populations in the laboratory. One of the best- the important qualitative features are already contained in studied examples is Lenski’s long-term evolution experiment Equation (40). (the LTEE), where 12 replicate populations of E. coli have

1258 B. H. Good and M. M. Desai Figure 4 The predicted fixation probability of a modifier allele when 6 26 Nsb ¼ 10 and e ¼ 10 : Grid points are colored according to the value of Npfix from Equation (38), and are capped at a maximum value of Np jlog2 fixj¼2 to maintain contrast. For comparison, the solid lines de- note 20 log-spaced contours that range from 1021 to 104: been propagated under constant conditions for more than 60;000 generations (Lenski et al. 1991, 2015). Mutator phe- notypes have fixedinsixofthe12populationsoverthe course of this experiment, and have typically increased the mutation rate by 100-fold. The rate of adaptation in the mutator populations increased less than twofold, sug- gesting that clonal interference plays an important role Figure 5 (Top) A vertical “slice” of Figure 4 for three different values U (Wiser et al. 2013). In one of the 12 replicates, the dynamics of . Symbols denote the results of forward time simulations with NU0 ¼ 5780; and sd ¼ 5sb; other parameters are the same as Figure 4. of mutation rate evolution have been studied in more detail. Solid lines denote theoretical predictions from Equation (38). (Bottom) Wielgoss et al. (2013) have shown that, soon after the fix- The scaled rate of adaptation as a function of the mutation rate for the ation of the mutator allele, an antimutator phenotype fixed, same set of parameters. Symbols denote the results of forward-time which lowered the mutation rate by a factor of approxi- simulations, the solid lines show the theoretical predictions obtained by solving Equations (31) and (32) numerically, and the dashed line shows mately two. The antimutator phenotype had two inde- 2 2 the asymptotic formula, v=s 2logðNsbÞ=logðsb=eUmÞ : pendent origins, suggesting that both it and the original b mutator allele were favored by selection. Given these find- ings, it is interesting to compare various explanations for account for these effects in a crude way by examining how U^ this behavior in the context of our theoretical analysis changes after a shift in Ub or sb: For example, epistasis could above. lead to a reduction in e,reflecting a diminished supply of One potential explanation forthese findings is suggested by beneficial mutations (or an increasing supply of deleterious the “overshooting” behavior illustrated in Figure 4 and Figure mutations) as the population adapts. Due to the presence of 5. Although the precise evolutionary parameters of these clonal interference, Equation (25) suggests that rather dras- populations are not known, Figure 5 shows that it is at least tic reductions in e are required to cause a twofold reduction in feasible to overshoot the optimum by a factor of 2 when the U^: Alternatively, diminishing-returns epistasis could reduce ^ population starts at a mutation rate of U0 0:01U: Note, the magnitude of sb; while leaving the beneficial mutation however, that Npfix is not necessarily large in the reverse di- rate unchanged. Equation (25) suggests that this has much rection, so it remains to be seen whether this effect could stronger effect on U^: Wiser et al. (2013) have recently pro- efficiently select for lower mutation rates on the rapid time- posed and fit a concrete model for how sb declines with fit- scales observed. ness in the LTEE. These estimates suggest that sb declines by Another potential explanation for the mutation-rate re- ≲20% in the 4000 generations that separate the mutator and versal, initially suggested by Wielgoss et al. (2013), is that antimutator alleles. This would seem to be too small to ac- long-term epistasis reduced the effective advantage of hitch- count for a ≳50% reduction in U^; although the magnitudes hiking later in the experiment, thereby shifting the location of are sufficiently close that a careful comparison with simula- U^ below its initial value. Our analysis above does not account tions is required. This would be an interesting avenue for for these effects directly, since it assumes a constant value of future work. e and sb: However, as long as any epistasis manifests itself on A third potential explanation is a direct fitness benefit timescales that are longer than the fixation time, we can still to either the mutator or antimutator allele. While such

Mutators in Rapidly Adapting Populations 1259 benefits are difficulttomeasuredirectly(duetothecon- 1994; Peck 1994). Our previous work on the effects of founding effects of the deleterious load), they can be in- deleterious mutations in adapting populations provides a corporated into our model in a straightforward way.We can detailed analysis of when these conditions will hold, and of simply replace Udðr 2 1Þ/Udðr 2 1Þ 2 sm in our formulae the effects of deleterious mutations on adaptation when for pfixðrÞ; where sm is the direct benefit of the modifier the purgeable assumption fails (Good and Desai 2014). allele. Since the effect of the deleterious load is typically Thisearlierworksuggeststhatdeleteriousmutationsare ^ of order U , sb; even a small direct benefit(sm , sb)can likely to be purgeable in most microbial evolution experi- shift the balance between hitchhiking and load in impor- ments, though recent experimental work hints that this tant ways. may not always be the case (McDonald et al. 2016). Here, we first summarize the conditions under which deleterious The evolution of mutation rates and the average rate of adaptation mutations are purgeable, and then describe the potential effects of deleterious mutations when this condition is Becausewehaveassumedthatdeleteriousmutationsare violated. purgeable, increasing U always increases the rate of adap- In our earlier work, we showed that deleterious mutations tation, even though these increases may be small in the with fitness cost sd will typically not fix provided that presence of clonal interference. Our results therefore sug- sd 1=Tc; where Tc qtest is the coalescence timescale gest that mutation rates will evolve toward a stable muta- (Good and Desai 2014). Since sb 1=Tc in all the regimes tion rate that is less than what would be optimal for the that we consider, this will be true provided that sd is not much N population (which, in this case, is technically Uopt ¼ ). smaller than sb: Deleterious mutations can also hinder the Of course, as we increase mutation rates, the assumption fixation of beneficial mutations that arise in less-fit back- that deleterious mutations are purgeable will eventually grounds. However, provided that sd Ud; deleterious vari- fail. Somewhere above this point, there will be an optimal ants will be rapidly eliminated from the population, and mutation rate Uopt , N that maximizes the rate of adapta- most genetic backgrounds will be free of deleterious fi 24 tion [see, e.g., Orr (2000)]. To con rm that the stable mu- mutations. Assuming typical values of Ud 10 and ; ^ 22 21 tation rate is indeed below Uopt we must verify that U is still sd 10 2 10 for microbial evolution experiments (Wloch in the purgeable deleterious mutations regime. For the ex- et al. 2001), this condition will often be met. ample in Figure 5, we can verify this directly, since the rate When deleterious mutations have small enough fitness of adaptation continues to increase as the mutation rate costs that they are not purgeable, our quantitative results passes through U^: This behavior will hold more generally all break down. However, sufficiently weakly deleterious provided that the typical cost of a deleterious mutation, sd; is mutations cannot affect mutator or antimutator dynamics ^ ^ much larger than U: Since U is typically smaller than sb in on the relevant timescales, so they are effectively neutral from the regimes that we consider, this will indeed be the case the point of view of mutation rate evolution. On the other whenever sd ≳ sb: hand, sufficiently strongly deleterious mutations are purge- In these cases, the stable mutation rate is below the optimal able. Thus, it is only some range of deleterious mutations mutation rate, which implies that the dynamics of the evolu- between these limits whose effects are both important to tionary process can sometimes favor changes in mutation rate mutation rate evolution and not described by our analysis. that slow the adaptation of the population. Antimutator Since these mutations are less deleterious than purgeable alleles can be favored even when their fixation will ultimately mutations, they must, by definition, be less unfavorable to reduce the overall rate of adaptation. Conversely, mutator mutator alleles (and less favorable to antimutators). Thus, alleles can be disfavored even when their fixation would have they do not affect the fate of a modifier of mutation rates as increased the rate of adaptation. strongly as a corresponding purgeable mutation would. This Although our results are limited to the purgeable regime, a suggests that we can qualitatively understand the effects of ^ similar distinction between U and Uopt may also apply even nonpurgeable deleterious mutations in adapting populations when deleterious mutations are no longer purgeable. We by weighting them less heavily than purgeable ones, so that cannot prove this conjecture in our present framework, but the total effective deleterious mutation rate is actually some- it is an interesting hypothesis for future work. what less than Ud: Of course, this is only a rough intuition. To analyze the effects of nonpurgeable deleterious mutations Limitations to our analysis more fully, we need to include them in our traveling wave We have made a number of key assumptions throughout our framework. Our earlier work explains how these mutations analysis. Most crucially, we have assumed that deleterious affect fðxÞ; wðxÞ; xc; and v, which provides a potential starting mutations are purgeable. For this to be true, two conditions point for these calculations (Good and Desai 2014). More must be met: the deleterious mutations cannot fix, nor generally, we may wish to consider the evolution of mutation can they affect the overall rate of adaptation by reducing rates in the limit where nonpurgeable mutations become so the fixation probabilities of beneficial mutations. This is important that the population no longer adapts, and instead a slightly stronger condition than the “ruby in the rough” Muller’s ratchet plays a crucial role. These are interesting but approximation used by earlier authors (Charlesworth complex directions for future work.

1260 B. H. Good and M. M. Desai In addition to these key limitations of our analysis, we have regime, Tfix logðsb=UbÞ=sb; and this condition becomes also made a number of more technical assumptions. For m sb=logðsb=UbÞNpfixðrÞ: This can sometimes be violated example, we have assumed that beneficial mutations all pro- for strongly beneficial mutator alleles with a large target size vide the same fitness benefit, sb: In reality, beneficial muta- (e.g., loss-of-function mutations in multiple genes). In these tions will have a range of different fitness effects, drawn from cases, our quantitative predictions become inaccurate, al- some distribution. However, earlier work has shown that in though the overall direction of selection [i.e., whether this case the evolutionary dynamics can often be summarized Npfix . 1orNpfix , 1] will remain unchanged. using a single effective beneficial fitness effect, and corre- Conclusions sponding effective beneficial mutation rate (Desai and Fisher 2007; Good et al. 2012; Fisher 2013). Thus our analysis Our analysis has explained how the interplay between can be applied to this situation using the appropriate effective beneficial and deleterious mutations in adapting popula- fi sb and Ub: tions creates indirect selection pressures on modi ers of In our analysis of clonal interference, we focused on a regime mutation rates, and we have shown how these indirect 2 in which 1 logðsb=UbÞ2logðNsbÞlog ðsb=UbÞ: The lat- selection pressures affect the fates of mutator and antimu- ter condition implies that clonal interference is not excep- tator alleles. Our analysis of the successional mutations tionally strong, and is often a good approximation for regime follows the logic of earlier work (Andre and Godelle microbial evolution experiments at wild-type mutation 2006; Desai and Fisher 2011), balancing the probability a rates. For some large effect mutators, we start to approach new beneficial mutation arises in a mutator, or antimuta- the boundary of this regime, and our quantitative expres- tor, background with the effects of the modifier allele on sionsbecomelessaccurate(seeFigure5).Inprinci- theaccumulationofdeleteriousload.Wehavealsostudied ple, we could extend our analysis to the “high-speed” rapidly adapting populations where clonal interference is 2 [2logðNsbÞlog ðsb=UbÞ]or“mutational-diffusion” re- widespread. Our approach to this question builds on the gimes [Ub sb] by using analogous solutions for fðxÞ and traveling wave framework we and others have recently wðxÞ derived by Fisher (2013) or Hallatschek (2011). We introduced (Neher et al. 2010; Hallatschek 2011; Neher leave this for future work. and Shraiman 2011; Good et al. 2012; Fisher 2013; Good Throughout our analysis, we have assumed that modifier and Desai 2014). In this framework, analyzing the fate of effect sizes can be large but not exceptionally so [e.g., in the an allele modifying mutation rate is similar in spirit to fi clonal interference regime, jlogðrÞj logðsb=UbÞ]. Since calculating the xation probability of any lineage: we must ; Ub sb; this includes many realistically large mutators of solve the same equation for wðxÞ except instead of a mu- order r 100; and, in practice, our expressions appear to tation changing the fitness x, it changes the mutation rate work reasonably well even when logðrÞlogðsb=UbÞ (see U. This general framework can also be applied to analyze Figure 3). For sufficiently large r, we may also encounter indirect selection pressures that act on other modifiers of situations where modifiers switch from one regime to an- the evolutionary process. For example, we could analyze other (e.g., from the clonal interference to successive the fate of a mutation that changes the distribution of fit- mutations regime, or from Ub sb to Ub . sb). Our quanti- ness effects of new mutations by solving for wðxÞ for an fi r : tative predictions for Npfix break down in all of these cases. allele that modi es ðsÞ Ouranalysisdoesnotneedtobe However, the evolutionarily stable mutation rates are de- limited to adapting populations, since the traveling wave fined by the behavior of Npfix in the local neighborhood of framework applies whenever interference selection is r 1; so the location of U^ is not influenced by this problem. widespread, even if the population is not adapting on av- If the population starts sufficiently far from U^; then erage. Of course, in practice, some modifiers and parameter the fixation probabilities of the first few Ui are not well- regimes will lead to equations for wðxÞ that are analytically described by our analysis, but we know that the mutation tractable, while others will not. Thus further work is needed rate must still approach U^ (in a possibly nonmonotonic to more fully understand both the limits and promise of manner). Eventually, the mutation rate will become close this approach. enough to U^ that our expressions start to apply, and the remaining steps of the mutation trajectory can be Acknowledgments predicted. Finally, we have assumed throughout that modifier mu- We thank Ivana Cvijovi´c for useful discussions and helpful tations are sufficiently rare that their fates are deter- comments on the manuscript. Simulations in this article mined independently. In other words, we have neglected were run on the Odyssey cluster supported by the Faculty clonal interference between different modifier mutations. of Arts and Sciences Division of Science Research Com- For deleterious or weakly beneficial modifiers, this will puting Group at Harvard University. This work was sup- often be a good approximation. However, for strongly ported in part by the Simons Foundation (grant 376196), beneficial modifiers, this requires that the establishment grant PHY 1313638 from the National Science Founda- time 1=NmpfixðrÞ between successive mutators is large tion, and grant GM-104239 from the National Institutes compared to the fixation time. In the clonal interference of Health.

Mutators in Rapidly Adapting Populations 1261 Literature Cited Good, B. H., I. M. Rouzine, D. J. Balick, O. Hallatschek, and M. M. Desai, 2012 Distribution of fixed beneficial mutations and the Andre, J.-B., and B. Godelle, 2006 The evolution of mutation rate rate of adaptation in asexual populations. Proc. Natl. Acad. Sci. in finite asexual populations. Genetics 172: 611–626. USA 109: 4950–4955. Barroso-Batista, J., A. Sousa, M. Lourenço, M.-L. Bergman, J. Good, B. H., A. M. Walczak, R. A. Neher, and M. M. Desai, Demengeot et al.,2014Thefirst steps of adaptation of 2014 Genetic diversity in the interference selection limit. PLoS Escherichia coli to the gut are dominated by soft sweeps. PLoS Genet. 10: e1004222. Genet. 10: e1004182. Hallatschek, O., 2011 The noisy edge of traveling waves. Proc. Bjorkholm, B., M. Sjolund, P. G. Falk, O. G. Berg, L. Engstrand et al., Natl. Acad. Sci. USA 108: 1783–1787. 2001 Mutation frequency and biological cost of antibiotic re- Ishii, K., H. Matsuda, Y. Iawasa, and A. Sasaki, 1989 Evolutionarily sistance in Helicobacter pylori. Proc. Natl. Acad. Sci. USA 98: stable mutation rate in a periodically changing environment. Ge- 14607–14612. netics 121: 163–174. Chao, L., and E. C. Cox, 1983 Competition between high and low James, A., and K. Jain, 2015 Fixation probability of rare nonmu- mutating strains of Escherichia coli. Evolution 37: 125–134. tator and evolution of mutation rates. Ecol. Evol. 6: 755–764. Chao, L., C. Vargas, B. B. Spear, and E. C. Cox, 1983 Transposable Johnson, T., 1999a The approach to mutation-selection balance elements as mutator genes in evolution. Nature 303: 633–635. in an infinite asexual population, and the evolution of mutation Charlesworth, B., 1994 The effect of background selection against rates. Proc. Biol. Sci. 266: 2389–2397. deleterious mutations on weakly selected, linked variants. Johnson, T., 1999b Beneficial mutations, hitchhiking and the evo- Genet. Res. 63: 213–227. lution of mutation rates in sexual populations. Genetics 151: Dawson, K. J., 1998 Evolutionarily stable mutation rates. J. Theor. 1621–1631. Biol. 194: 147–157. Kessler, D. A., and H. Levine, 1998 Mutator dynamics on a smooth Dawson, K. J., 1999 The dynamics of infinitesimally rare alleles, evolutionary landscape. Phys. Rev. Lett. 80: 2012–2015. applied to the evolution of mutation rates and the expression of Kimura, M., 1967 On the evolutionary adjustment of spontaneous deleterious mutations. Theor. Popul. Biol. 55: 1–22. mutation rates. Genet. Res. 9: 23–34. del Campo, R., M. I. Morosini, E. G. G. de la Pedrosa, A. Fenoll, C. Labat, F., O. Pradillon, L. Garry, M. Peuchmaur, B. Fantin et al., Munoz-Almagro et al., 2005 Population structure, antimicro- 2005 Mutator phenotype confers advantage in Escherichia coli bial resistance, and mutation frequencies of Streptococcus pneu- chronic urinary tract infection pathogenesis. FEMS Immunol. moniae isolates from cystic fibrosis patients. J. Clin. Microbiol. Med. Microbiol. 44: 317–321. 43: 2207–2214. LeClerc, J. E., B. G. Li, W. L. Payne, and T. A. Cebula, 1996 High Denamur, E., S. Bonacorsi, A. Giraud, P. Duriez, F. Hilali et al., mutation frequencies among Escherichia coli and Salmonella 2002 High frequency of mutator strains among human uropa- pathogens. Science 274: 1208–1211. thogenic Escherichia coli isolates. J. Bacteriol. 184: 605–609. Lee, H., E. Popodi, H. Tanga, and P. L. Foster, 2012 Rate and Desai, M. M., and D. S. Fisher, 2007 Beneficial mutation selection molecular spectrum of spontaneous mutations in the bacterium balance and the effect of genetic linkage on positive selection. Escherichia coli as determined by whole-genome sequencing. Genetics 176: 1759–1798. Proc. Natl. Acad. Sci. USA 109: E2774–E2783. Desai, M. M., and D. S. Fisher, 2011 The balance between muta- Leigh, E. G., 1970 Natural selection and mutability. Am. Nat. tors and nonmutators in asexual populations. Genetics 188: 104: 301. 997–1014. Leigh, E. G., 1973 The evolution of mutation rates. Genetics 73: Desai, M. M., D. S. Fisher, and A. W. Murray, 2007 The speed of 1–18. evolution and the maintenance of variation in asexual popula- Lenski, R. E., M. R. Rose, S. C. Simpson, and S. C. Tadler, tions. Curr. Biol. 17: 385–394. 1991 Long-term experimental evolution in Escherichia coli. i. Desai, M. M., A. M. Walczak, and D. S. Fisher, 2013 Genetic di- Adaptation and divergence during 2,000 generations. Am. Nat. versity and the structure of genealogies in rapidly adapting pop- 138: 1315–1341. ulations. Genetics 193: 565–585. Lenski, R. E., M. J. Wiser, N. Ribeck, Z. D. Blount, J. R. Nahum et al., Fisher, D. S., 2007 Evolutionary dynamics, pp. 395–446. in Com- 2015 Sustained fitness gains and variability in fitness trajecto- plex Systems, Vol. 85, edited by M. M. Jean-Philippe Bouchaud, ries in the long-term evolution experiment with Escherichia coli. and J. Dalibard. Elsevier, Les Houches. Proc. Biol. Sci. 282: 20152292. Fisher, D. S., 2013 Asexual evolution waves: fluctuations and uni- Liberman, U., and M. W. Feldman, 1986 Modifiers of mutation rate: versality. J. Stat. Mech. 2013: P01011. a general reduction principle. Theor. Popul. Biol. 30: 125–142. Gardiner, C., 1985 Handbook of Stochastic Methods. Springer, Lynch, M., 2008 The cellular, developmental and population- New York. genetic determinants of mutation-rate evolution. Genetics Gentile, C. F., S.-C. Yu, S. A. Serrano, P. J. Gerrish, and P. D. 180: 933–943. Sniegowski, 2011 Competition between high- and higher- Lynch, M., 2011 The lower bound to the evolution of mutation mutating strains of Escherichia coli.Biol.Lett.7:422–424. rates. Genome Biol. Evol. 3: 1107–1118. Gillespie, J. H., 1981 Mutation modification in a random environ- Matic, I., M. Radman, F. Taddei, B. Picard, C. Doit et al., ment. Evolution 35: 468–476. 1997 Highly variable mutation rates in commensal and path- Giraud, A., I. Matic, O. Tenaillon, A. Clara, M. Radman et al., ogenic Escherichia coli. Science 277: 1833–1834. 2001 Costs and benefits of high mutation rates: adaptive evo- McDonald, M. J., Y. Y. Hsieh, Y. H. Yu, S. L. Chang, and J. Y. Leu, lution of bacteria in the mouse gut. Science 291: 2606–2608. 2012 The evolution of low mutation rates in experimental Giraud, A., I. Matic, M. Radman, M. Fons, and F. Taddei, mutator populations of Saccharomyces cerevisiae. Curr. Biol. 2002 Mutator bacteria as a risk factor in treatment of infec- 22: 1235–1240. tious diseases. Antimicrob. Agents Chemother. 46: 863–865. McDonald, M. J., D. P. Rice, and M. M. Desai, 2016 Sex speeds Good, B. H., and M. M. Desai, 2014 Deleterious passengers in adaptation by altering the dynamics of molecular evolution. adapting populations. Genetics 198: 1183–1208. Nature 531: 233–236. Good, B. H., and M. M. Desai, 2015 The impact of macroscopic Neher, R., and B. I. Shraiman, 2011 Genetic draft and quasi- epistasis on long-term evolutionary dynamics. Genetics 199: neutrality in large facultatively sexual populations. Genetics 177–190. 188: 975–976.

1262 B. H. Good and M. M. Desai Neher, R., and B. I. Shraiman, 2012 Fluctuations of fitness distribu- Tenaillon, O., H. L. Nagard, B. Godelle, and F. Taddei, tions and the rate of Muller’s ratchet. Genetics 191: 1283–1293. 2000 Mutators and sex in bacteria: conflict between adaptive Neher, R., B. I. Shraiman, and D. S. Fisher, 2010 Rate of adapta- strategies. Proc. Natl. Acad. Sci. USA 97: 10465–10470. tion in large sexual populations. Genetics 184: 467–481. Thompson, D. A., M. M. Desai, and A. W. Murray, 2006 Ploidy Neher, R. A., T. A. Kessinger, and B. I. Shraiman, 2013 Coales- controls the success of mutators and nature of mutations during cence and genetic diversity in sexual populations under selec- budding yeast evolution. Curr. Biol. 16: 1581–1590. tion. Proc. Natl. Acad. Sci. USA 110: 15836–15841. Travis, J., and E. Travis, 2002 Mutator dynamics in fluctuating Notley-McRobb, L., S. Seeto, and T. Ferenci, 2002 Enrichment environments. Proc. Biol. Sci. 269: 591–597. and elimination of mutY mutators in Escherichia coli popula- Tröbner, W., and R. Piechocki, 1981 Competition growth between tions. Genetics 162: 1055–1062. Escherichia coli mutL and mut+ in continuously growing cul- Oliver, A., R. Canton, P. Campo, F. Baquero, and J. Blazquez, tures. Z. Allg. Mikrobiol. 21: 347–349. 2000 High frequency of hypermutable Pseudomonas aerugi- Tröbner, W., and R. Piechocki, 1984 Selection against hypermu- nosa in cystic fibrosis lung infection. Science 288: 1251–1253. tability in Escherichia coli during long term evolution. Mol. Gen. Orr, H. A., 2000 The rate of adaptation in asexuals. Genetics 155: Genet. 198: 177–178. 961–968. Tsimring, L., H. Levine, and D. Kessler, 1996 RNA virus evolution Painter, P. R., 1975 Mutator genes and selection for the mutation via a fitness-space model. Phys. Rev. Lett. 90: 088103. rate in bacteria. Genetics 79: 649–660. Turrientes, M.-C., F. Baquero, B. R. Levin, J.-L. Martinez, A. Ripoll Pal, C., M. D. Macia, A. Oliver, I. Schachar, and A. Buckling, et al., 2013 Normal mutation rate variants arise in a mutator 2007 Coevolution with viruses drives the evolution of bacte- (mut s) Escherichia coli population. PLoS One 8: e72963. rial mutation rates. Nature 450: 1079–1081. Voordeckers, K., J. Kominek, A. Das, A. Espinosa-Cantú, D. D. Peck, J. R., 1994 A ruby in the rubbish. Genetics 137: 597–606. Maeyer et al., 2015 Adaptation to high ethanol reveals com- Perfeito, L., L. Fernandes, C. Mota, and I. Gordo, 2007 Adaptive mu- plex evolutionary pathways. PLoS Genet. 11: e1005635. tations in bacteria: high rate and small effects. Science 317: 813. Watson, M. E., J. L. Burns, and A. L. Smith, 2004 Hypermutable Prunier, A. L., B. Malbruny, M. Laurans, J. Brouard, J. F. Duhamel haemophilus influenzae with mutations in muts are found in et al., 2003 High rate of macrolide resistance in Staphylococcus cystic fibrosis sputum. Microbiol. Sgm. 150: 2947–2958. aureus strains from patients with cystic fibrosis reveals high pro- Weissman, D., and O. Hallatschek, 2014 The rate of adaptation in portions of hypermutable strains. J. Infect. Dis. 187: 1709–1716. large sexual populations with linear chromosomes. Genetics Richardson, A. R., Z. Yu, T. Popovic, and I. Stojiljkovic, 196: 1167–1183. 2002 Mutator clones of neisseria meningitidis in epidemic se- Weissman, D. W., M. M. Desai, D. S. Fisher, and M. W. Feldman, rogroup a disease. Proc. Natl. Acad. Sci. USA 99: 6103–6107. 2009 The rate at which asexual populations cross fitness val- Rouzine, I. M., J. Wakeley, and J. M. Coffin, 2003 The solitary wave leys. Theor. Popul. Biol. 75: 286–300. of asexual evolution. Proc. Natl. Acad. Sci. USA 100: 587–592. Wielgoss, S., J. E. Barrick, O. Tenaillon, S. Cruveiller, B. Chane- Shaver, A. C., P. G. Dombrowski, J. Y. Sweeney, T. Treis, R. M. Woon-Ming et al., 2011 Mutation rate inferred from synony- Zappala et al., 2002 Fitness evolution and the rise of mutator mous substitutions in a long-term evolution experiment with alleles in experimental Escherichia coli populations. Genetics Escherichia coli. G3 (Bethesda) 1: 183–186. 162: 557–566. Wielgoss, S., J. E. Barrick, O. Tenaillon, M. J. Wiser, J. Dittmar Sniegowski, P. D., P. J. Gerrish, and R. E. Lenski, 1997 Evolution et al., 2013 Mutation rate dynamics in a bacterial population of high mutation rates in experimental populations of E. coli. reflect tension between adaptation and . Proc. Natl. Nature 387: 703–705. Acad. Sci. USA 110: 222–227. Soderberg, R. J., and O. G. Berg, 2011 Kick-starting the ratchet: Wiser, M. J., N. Ribeck, and R. E. Lenski, 2013 Long-term dynam- the fate of mutators in asexual populations. Genetics 197: ics of adaptation in asexual populations. Science 342: 1364– 1129–1137. 1367. Taddei, F., M. Radman, J. Maynard-Smith, B. Toupance, P. H. Wloch, D. M., K. Szafraniec, R. H. Borts, and R. Korona, Gouyon et al., 1997 Role of mutator alleles in adaptive evolu- 2001 Direct estimate of the mutation rate and the distribution tion. Nature 387: 700–702. of fitness effects in the yeast Saccharomyces cerevisiae. Genetics Tanaka, M. M., C. T. Bergstrom, and B. R. Levin, 2003 The evo- 159: 441–452. lution of mutator genes in bacterial populations: the roles of Wylie, C. S., C.-M. Ghim, D. Kessler, and H. Levine, 2009 The environmental change and timing. Genetics 164: 843–854. fixation probability of rare mutators in finite asexual popula- Tenaillon,O.,B.Toupance,H.L.Nagard,F.Taddei,andB.Godelle, tions. Genetics 181: 1595–1612. 1999 Mutators, population size, adaptive landscape, and the ad- aptation of asexual populations of bacteria. Genetics 152: 485–493. Communicating editor: S. I. Wright

Mutators in Rapidly Adapting Populations 1263 Appendix A: Exact Solution of Equation (3) In this section, we evaluate the integral Z N R 2 t l 9 l 9 9 ½ 0ðt Þþ mðt Þdt Npfix ¼ lmðtÞe 0 dt ; (A1) 0 where l ¼ 2NU s ½1 2 f ðtÞ and l ðtÞ¼2NU s rf ðtÞ; and f ðtÞ satisfies the Langevin dynamics 0 b b m m b b m rmffiffiffiffiffi @f f m ¼ mhðtÞ; (A2) @t N with fmð0Þ¼1=N: Using integration by parts, we can rewrite this integral as  Z R  N t r dt 2 t 2r21 9 9 t t fmðt Þ dt Npfix ¼ 1 2 e est e est 0 ; (A3) r 2 1 0 test t = : fi where we have used theR short-hand est ¼ 1 2NUbsb To evaluate this integral, we must rst solve for the generating function t Hð y; tÞ for the weight fmðt9Þdt9; which is defined by 0 R 2 t 9 9 y fmðt Þ dt Hð y; tÞ¼ e 0 : (A4)

This can be done using techniques described in Weissman et al. (2009). Briefly, we first introduce the joint generating function, R 2 2 t 9 9 zfmðtÞ y fmðt Þ dt Hðz; y; tÞ ¼ e 0 : (A5)

By Taylor expanding Hðz; y; t þ dtÞ and using the Langevin dynamics in Equation (A2), we can show that Hðz; y; tÞ obeys the partial differential equation   @H z2 @H ¼ y 2 ; (A6) @t 2N @z which can be solved using the method of characteristics. The characteristic curves satisfy @y @z z2 @H ¼ 0 ; ¼ 2 y ; ¼ 0 ; (A7) @s @s 2N @s and the solution is given by " # pffiffiffiffiffiffiffiffiffi   2Ny t pffiffiffiffiffiffiffiffiffi z Hðz; y; tÞ¼1 2 tanh 2Ny þ tanh1 pffiffiffiffiffiffiffiffiffi : (A8) N 2N 2Ny R t The marginal generating function for fmðt9Þ dt9 can then be obtained by setting z ¼ 0: 0 pffiffiffiffiffiffiffiffiffi   2Ny t pffiffiffiffiffiffiffiffiffi Hð y; tÞ¼1 2 tanh 2Ny : (A9) N 2N

Substituting this expression into the integral in Equation (A3), we find that

Z sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ! N t 2 2z 2N estðr 1Þ Npfix ¼ r e tanh z dz; (A10) 0 testðr 2 1Þ 2N 2  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3 t 2 C 1 N 2 C N 2 2 estðr 1Þ 6 2 þ 8t ðr 2 1Þ 8t ðr 2 1Þ N 7 ¼ r4 est est 5; (A11) testðr 2 1Þ N where CðzÞ is the digamma function. Asymptotic expansions for small and large arguments yield

1264 B. H. Good and M. M. Desai 8 "  # > t ðr 2 1Þ > r 1 2 O est if t r N; <> N est Npfix " rffiffiffiffiffiffiffiffiffi !# (A12) > pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi > N :> Nr=test 1 2 O if testr N: testr

Appendix B: Derivation of Equation (38)

To obtain an expression for NpfixðrÞ; we must evaluate the integral

Z Z Z À Á À Á À Á À Á À Á À Á ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ pfixðrÞ¼ f x wm x dx ¼ f x wm x 2 Udðr 2 1Þ dx ¼ f x þ Udðr 2 1Þ wm x dx; (B1) using the expressions for fðx~Þ and w~ mðx~Þ in Equations (29) and (34) in the main text. This is straightforward, although the algebra is somewhat tedious. Since fðx~Þ and w~ mðx~Þ are piecewise functions, the form of the integrand will depend on whether xc 2 Udðr 2 1Þ is greater or less than xcm and xcm 2 sb: We first consider the case where xc 2 Udðr 2 1Þ . xcm: In this case, the mutator interference threshold is reached before the nose, so that

2 2 Z ðÞxþUdðr 1Þ Z x2 h i xcm 2 22 2 xc 2 e 2v x xcm e 2v p ðrÞ pffiffiffiffiffiffiffiffiffi 2x e 2v dx~ þ pffiffiffiffiffiffiffiffiffi 2 x~ 2 U ðr 2 1Þ dx~: (B2) fix p cm p d xcm2sb 2 v xcmþUdðr21Þ 2 v pffiffiffiffiffiffiffiffiffi 2 2pv xc After multiplying both sides by N ¼ e2v ; we obtain 2xcsb

2 Z 22 2 U r21 2 sb 2 xc xcm2ð dð ÞÞ 2xcmUdðr 1Þ xcm Udðr 1Þz NpfixðrÞe 2v 2v v e v dz xcsb 0 2 xc Z Z z2 xc xc 2 e2v 2z2 e 2v þ ze 2v dz22NU ðr 2 1Þ pffiffiffiffiffiffiffiffiffi dz: (B3) d p xcsb xcmþUdðr21Þ xcmþUdðr21Þ 2 v

v Evaluating the remaining integrals and substituting xcm ¼ xc 2 logðrÞ; we find that h i h sib 2 3 2 2 2 U s ðr21Þ xc Udsbðr 1Þ v Udsbðr 1Þ d b logðrÞ2 2 logðrÞ2 v 2 sb v 2s2 v 6e 17 NpfixðrÞe b 4 5 Udsbðr 2 1Þ 2 h i h iv 3 2 U s r21 U s r21 xc logðrÞ2 d bð Þ 2 v logðrÞ2 d bð Þ v 4 sb v 2s2 v 5 þ e b 2 1 xcsb .   .  À v pffiffiffiffiffi pffiffiffiffiffi Erfc x 2 logðrÞþU ðr 2 1Þ 2v 2 Erfc x 2v c s d c 22NU ðr 2 1Þ b ; (B4) d 2 R N 2 2 p2ffiffiffi t : where ErfcðzÞ¼ p z e dt The last two terms are only relevant for extremely high deleterious mutation rates where Udð1 2 rÞxc: After neglecting these terms, we obtain Equation (11) in the main text. Now we consider the case where xc 2 Udðr 2 1Þ , xcm: In this case, the edge of the fitness distribution is reached before the mutator interference threshold. If xc 2 Udðr 2 1Þ , xcm 2 sb; then the edge of the fitness distribution never even makes it to the ~ ~ region where wmðxÞ is positive, and pfixðrÞ0: On the other hand, if xc 2 Udðr 2 1Þ lies between xcm and xcm 2 sb; then

2 2 Z ðÞxþUdðr 1Þ xc2Udðr21Þ 2 22 2 e 2v x xcm ffiffiffiffiffiffiffiffiffi 2v ~ pfixðrÞ¼ p 2xcme dx: (B5) x 2s 2pv pffiffiffiffiffiffiffiffiffi cm b 2 2pv xc Multiplying by N ¼ e2v, and evaluating the integrals, we obtain 2xcsb

Mutators in Rapidly Adapting Populations 1265 h i h i 2 3 2 2 2 2 U ðr21Þ xc Udsbðr 1Þ Udsbðr 1Þ v Udsbðr 1Þ 2 d 2 2 2 logðrÞ2 þ 2 logðrÞ 2 2 v ðxc Udðr 1Þ xcmþsbÞ sb v v 2s2 v 61 e 7 NpfixðrÞ¼e b 4 5: (B6) Udsbðr 2 1Þ v

In practice, the differences between Equations (B4) and (B6) do not become very pronounced until NpfixðrÞ is already quite small. For simplicity, we will therefore use Equation (B4) for the full range of r.

Appendix C: Forward-Time Simulations We validate several of our key approximations by comparing our predictions with the results of forward-time simulations similar to the standard Wright-Fisher model. In these simulations, the population is divided into fitness classes depending on the number of beneficial mutations (nb), and the number of deleterious mutations (nd), that each individual possesses. The simulation starts with a homogeneous population of N individuals with nb ¼ 0 and nd ¼ 0: To form the next generation, each individual produces a Poisson number of identical offspring with mean Ctð1 þ sbnb 2 sdndÞð1 2 Ud 2 UbÞ; a Poisson number of beneficial offspring (nb/nb þ 1) with mean Ctð1 þ sbnb 2 sdndÞUb; and a Poisson number of deleterious offspring (nd/nd þ 1) with mean Ctð1 þ sbnb 2 sdndÞUd; where Ct is a constant recalculated at each generation to ensure that the expected number of offspring for the entire population is N. Starting from the initial population, all simulations are allowed to “burn-in” for Dt ¼ 5 3 104 generations before any measurements are made. To measure the fixation probability of a modifier allele, we modify this basic algorithm so that new modifier offspring are produced from wild-type individuals at rate m. These modifier lineages reproduce the same way as wild-type individuals except with Ub/rUb and Ud/rUd: Further modifications to the mutation rate, or reversion to the wild type, are not allowed. To minimize the effects of initial conditions, m is artificially fixed at zero until the burn-in period has elapsed. We then record the number of generations, T, between the end of the burn-in period, and the fixation of the modifier phenotype. In the limit that m/0; this is related to the fixation probability through the relation

1 Np ðrÞ¼ : (C1) fix mT

To ensure that m is chosen to be small enough for Equation (C1) to apply, we repeat this process M ¼ 60 times, with a sequence fi m ; m ; ...; m ; fi ; ; ...; : of modi er mutation rates 1 2 M and a sequence of xation times T1 T2 TM The mutation rate at step i is chosen 4 based on the previous T 2 ; so that the predicted value of T is 10 generations: i 1 i m 2 T 2 2 m ¼ min i 1 i 1; 10 2 : (C2) i 104

The value 104 is chosen because, for the parameters considered here, 104 1 logðs =U Þ; and the mutation rates are capped at sb b b 1022 to minimize the correlated effects of Muller’s ratchet (Neher and Shraiman 2012). The sequence is started with m 24; “ ” ’ fi 1 ¼ 10 and is allowed to burn-in for 10 iterations before Ti s are recorded. The xation probability is calculated from the maximum likelihood estimator, b 1 N p ðrÞ ¼ P : (C3) fix 60 m i¼10 iTi

A copy of our implementation in Python is available at https://github.com/benjaminhgood/mutator_simulations.

1266 B. H. Good and M. M. Desai