| PERSPECTIVES

Developmental Bias and : A Regulatory Network Perspective

Tobias Uller,*,1 Armin P. Moczek,† Richard A. Watson,‡ Paul M. Brakefield,§ and Kevin N. Laland** *Department of Biology, Lund University, 22362, Sweden, †Department of Biology, Indiana University, Bloomington, Indiana 47405, ‡Department of Computer Science, Institute for Life Sciences, Southampton SO17 1BJ, UK, §Department of Zoology, University of Cambridge, CB2 3EJ, UK, and **School of Biology, University of St. Andrews, Fife KY16 9TF, UK

ABSTRACT Phenotypic variation is generated by the processes of development, with some variants arising more readily than others—a phenomenon known as “developmental bias.” Developmental bias and have often been portrayed as alternative explanations, but this is a false dichotomy: developmental bias can evolve through natural selection, and bias and selection jointly influence phenotypic evolution. Here, we briefly review the evidence for developmental bias and illustrate how it is studied empirically. We describe recent theory on regulatory networks that explains why the influence of genetic and environmental perturbation on is typically not uniform, and may even be biased toward adaptive phenotypic variation. We show how bias produced by developmental processes constitutes an evolving property able to impose direction on adaptive evolution and influence patterns of taxonomic and phenotypic diversity. Taking these considerations together, we argue that it is not sufficient to accommodate de- velopmental bias into evolutionary theory merely as a constraint on evolutionary . The influence of natural selection in shaping developmental bias, and conversely, the influence of developmental bias in shaping subsequent opportunities for adaptation, requires mechanistic models of development to be expanded and incorporated into evolutionary theory. A regulatory network perspective on phenotypic evolution thus helps to integrate the generation of phenotypic variation with natural selection, leaving better placed to explain how organisms adapt and diversify.

KEYWORDS development; constraint; ; ; facilitated variation; developmental bias

HE extraordinary diversity and adaptive fitoforgan- Although the diversity of life may give the impression that Tisms evolving under natural selection depends fun- natural selection can produce any form, it is well-recognized damentally on the generation of heritable phenotypic and uncontroversial that not all phenotypic variants are variation. Phenotypes are the result of causal interactions possible or even likely to be generated (Darwin 1859; at multiple levels of biological organization, including Waddington 1957; Maynard-Smith et al. 1985). The bias im- genes, cells, tissues, and organisms and their environ- posed on the distribution of phenotypic variation, arising ments. Given the complexity of these interactions, it is from the structure, character, composition, or dynamics of usually not obvious how developmental and physiological the developmental system, relative to the assumption of iso- systems will respond to perturbations, such as when genes tropic variation, is known as developmental bias1 (Maynard- mutate or environments change. Yet, without knowledge Smith et al. 1985; Arthur 2004; Wilkins 2007). The concept of the nature of phenotypic variability,an understanding of of developmental bias2 thus captures the observation that how and why evolution unfolds in the manner that it does is perturbation (e.g., , environmental change) to biolog- woefully incomplete. Natural selection cannot work with ical systems will tend to produce some variants more readily, imaginary phenotypes, only those realized by developmen- or with higher probability than others. Only at the extreme is tal systems. this manifest as the complete inability to produce a trait. The organization of a biological system is a product of its Copyright © 2018 by the Genetics Society of America evolution. Both developmental systems that produce unbi- doi: https://doi.org/10.1534/genetics.118.300995 ased patterns of phenotypic variation and those that produce Manuscript received November 28, 2017; accepted for publication April 19, 2018. 1Corresponding author: Department of Biology, Sölvegatan 37, 223 62 Lund, Sweden. bias need an evolutionary explanation (Salazar-Ciudad 2006, Email: [email protected]. 2008).Thepropensitytovaryinresponsetoparticular

Genetics, Vol. 209, 949–966 August 2018 949 Box 1 Methods for detecting developmental bias As natural selection is expected to remove variation, studies of standing phenotypic variation in a population, , or higher taxa provides an unsatisfactory method to demonstrate bias. To establish developmental bias, researchers must study the propensity for developmental systems to vary (their variability) rather than the observed state of variation (Wagner and Altenberg 1996). Much of what we have learnt of developmental bias comes from detailed experimental studies of development that reveal causal dependencies producing correlated changes in phenotypes, sometimes allowing for the prediction of phenotypic form across multiple species. For example, decades of research have revealed how the development of the limb skeleton is regulated (Hall 2015), which makes it possible to explain and predict correlated changes in digit length and the ordered loss of digits over evolutionary time (e.g., Alberch and Gale 1985; Kavanagh et al. 2013). A more quantitative approach is to study the distribution of phenotypic variation caused by genetic or environ- mental perturbation. Experimental evolution (e.g., McDonald et al. 2009) and mutation accumulation lines (e.g., Houle et al. 2017) can establish if random mutation produces some phenotypes more frequently than others. Furthermore, gene- editing tools make it possible to study the effects of change to particular genes or regulatory elements (Nakamura et al. 2016). Individuals can be exposed to stress or novel environmental conditions to determine whether developmental systems produce some phenotypes more frequently than others (Badyaev 2009). Sometimes it is possible to represent developmental processes mathematically, which makes it possible to study variability in silico (Salazar-Ciudad and Jernvall 2010), and to use computational modeling to predict phenotypic variation in nature (e.g.,Kavanaghet al. 2007). As illustrated in the main text, some well-understood systems have been studied from several of these perspectives. genetic or environmental inputs can be under natural selec- Fortunately, recent methodological advances that afford tion (e.g., McNamara et al. 2016). It is less obvious, however, more detailed analyses of how organisms develop are shed- if and how fitness differences can explain phenotypic bias in ding light on how bias can arise and revealing its prevalence in response to nondirected (i.e., random) genetic mutation or nature (Box 1; Figure 1). For example, the regulation of the environments that have not been experienced in the recent tetrapod limb creates developmental bias in the number and evolutionary history. That phenotypic variation is unbiased distribution of digits, limbs, and segments (Alberch and Gale has therefore probably been the default assumption in evo- 1985; Wake 1991), and in the proportion of skeletal parts lutionary theory. Here we explain why this assumption is (Sanger et al. 2011; Kavanagh et al. 2013). Interactions be- likely to be unfounded. We show how mechanistic models tween the components of developmental systems also bias can reveal the influence of selection in shaping developmen- relationships between the size, shape, and position of struc- tal bias, and conversely, how developmental bias can shape tural and pigment coloration of insect wings (Brakefield and subsequent evolution. This body of theory suggests that de- Roskam 2006; Prud’homme et al. 2006), the shape of beaks velopmental bias is not only likely to be widespread, but that (Campas et al. 2010; Fritz et al. 2014), the positioning of it may also contribute to adaptation and diversification. We cephalic horns in scarab beetles (Busey et al. 2016), and end by illustrating how these predictions can be tested by flower morphology (Wessinger and Hileman 2016). combining empirical studies of developmental processes with Tooth morphology in mammals provides a particularly comparative analyses of evolutionary diversification. compelling example of how developmental studies can be combined with computational analyses to demonstrate bias. Salazar-Ciudad and Jernvall (2010) integrated molecular de- Evidence for Developmental Bias tails of the gene network underlying molar development in In a classic discussion of developmental bias, Raup (1966) mice with biomechanical properties of cells to build a com- showed that only a comparably small proportion of all possi- putational model of tooth development. Their models were ble snail shell shapes was realized in nature, and suggested able to reproduce accurately variation in teeth morphology that this was partly explained by the mechanics of growth observed within species (Salazar-Ciudad and Jernvall 2010), [see also McGhee (2007) and Brakefield (2008)]. However, it predict morphological patterns both across species and in is not possible to assess the role of developmental bias solely teeth cultivated in vitro (Kavanagh et al. 2007; Harjunmaa from an absence of forms in nature since such absence is also et al. 2014), and even retrieve ancestral character states predicted to arise if the evolutionary process has not yet had (Harjunmaa et al. 2012). sufficient time to explore all options, or through natural se- Developmental bias can also be studied by examining how lection, which restricts phenotypes to regions of phenotypic traits are affected by genetic mutation. Such studies reveal space that have adaptive value. Other approaches to identi- that when phenotypic effects do occur,random mutation often fying bias (e.g., genetic correlations between traits; Maynard- produces nonrandom distributions of phenotypes. For exam- Smith et al. 1985) have also proven inconclusive, which for ple, Braendle et al. (2010) conducted a thorough quantifica- many years left the prevalence and significance of develop- tion of the phenotypic variability of the vulval developmental mental bias difficult to ascertain. system across mutation accumulation lines of two species of

950 T. Uller et al. Figure 1 Compelling examples of devel- opmental bias and its evolutionary effect in animals. (A) By combining experiments in vivo and in vitro, comparative analyses, and mathematical modeling, researchers have shown that the evolutionary diver- sity in tooth morphology among mam- mals is shaped by the mechanism by which teeth develop. Pictured is the skull of a crabeater seal, Lobodon carcino- phaga. (B) The oral and pharyngeal jaws of cichlid fishes are putative examples of how a bias caused by plasticity, itself pos- sibly favored by selection, can feed back to facilitate adaptive divergence and con- vergence in independently evolving line- ages. (C) In Drosophila, the phenotypic divergence between species in wing shape is aligned with the phenotypic bias asso- ciated with random mutation, one expla- nation for which is that developmental bias coevolves with phenotypic diver- gence. (D) Artificial selection on the size and color of Mycalesine butterfly eye spots demonstrates the effects of bias misaligned or aligned with the direction of selection. Photo credits: (A) Panther Media GmbH, Alamy Stock Photo; (B) Kevin Parsons; (C) Martin Hauser Phycus, CC-BY-3.0-DE; (D) Saenko et al., BMC Biology 2010 8:111, CC-BY-2.0.

Caenorhabditis nematodes. The results demonstrated that Developmental Bias Is More Than Constraint spontaneous produce bias, with some phenotypic To the extent that the evolutionary biology literature con- variants being common and others rare or absent. The fea- siders bias, these are most commonly thought to be con- tures of the vulva that were most affected by new mutations straints: features of organisms that hinder, or even prevent, also tended to show a greater variation in the stock population, populations from evolving adaptively (Maynard-Smith et al. suggesting that this bias occurs in nature. Similarly, although 1985; Futuyma 2015). “Constraint” implies that some re- virtually all dimensions of the Drosophila wing are variable in gions of phenotypic space that are adaptive are not populated nature and in mutation accumulation lines (Scharloo 1970; by the phenotypic variation that arises in development. An Mezey and Houle 2005; Houle and Fierst 2013), mutations oft-cited example is the evolution of the mammalian neck disproportionately cause covariation among parts of the wing (e.g., Galis 1999). In contrast to birds and reptiles, elongation such that some shapes are more readily produced than others of the mammalian neck has exclusively taken place by mak- (Klingenberg and Zaklan 2000; Houle et al. 2017). In plants, ing the vertebrae larger rather than by adding vertebrae, as chemical mutagenesis has been shown to induce pheno- seen, for instance, in long-necked plesiosaurs. Viewed from typic variants with a biased covariance structure between an engineering or design perspective, vertebrae number fl plant growth, owering and seed set in Arabidopsis thaliana likely constrains the evolution of long, slender, and maneu- (Camara and Pigliucci 1999; Camara et al. 2000). verable necks in mammals. This absence of variants with Although development is often buffered against environ- additional neck vertebrae is apparently because mutations mental stress, environmental conditions can also profoundly that modify the number of cervical vertebrae disrupt funda- affect phenotypic variation and covariation, and environmen- mental features of the mammalian body plan (Galis et al. tally induced developmental bias is manifest in diverse taxa 2006). The uniform selection against those variants is not and contexts. For example, jaw morphology in vertebrates itself the constraint; rather, the mammalian phenotypic responds in characteristic ways to diet (e.g., Gomez-Mestre space is biased in part because mammalian developmental and Buchholz 2006; Young and Badyaey 2010; Young et al. biology struggles to produce variants with more than seven 2010; Muschick et al. 2011; Scott et al. 2014). More gener- vertebrae that also preserve the remainder of the body plan. ally, as witnessed in the house finch, stress-induced pheno- Suchformscouldbefavoredbyselection,weretheyto typic variation can be directional, channeled by existing appear. developmental pathways, and integrated across morpholog- While the term “developmental bias” is inclusive of devel- ical, endocrinal, and behavioral systems (Badyaev 2005, opmental constraint, it goes beyond it, as do its evolutionary 2009). Both genetic mutation and environmental stress con- implications. A categorical distinction between “what is pos- tribute to the developmental bias observed in congenital ab- sible” (i.e., no constraints, selection has a free reign) and normalities, where large and highly nonadaptive phenotypic “what is not possible” (constraints operating) neglects that variants have been shown to share structural regularities bias within the “what is possible” region can significantly across distantly related species (Alberch 1989). shape how and why evolution unfolds the way that it does.

Developmental Bias and Evolution 951 First, developmental systems can alter the ratio or proportion that bias contributes to evolvability, by which we mean the of variation that occurs on one phenotypic dimension relative capacity for a lineage to undergo adaptive evolution. This to another (Figure 2). Since altering the ratio of variability in phenomenon is partly captured by existing theoretical models different phenotypic dimensions can influence the direction (e.g., Jones et al. 2007, 2014; Pavlicev et al. 2011). However, of evolutionary change (Arnold et al. 2001), developmental a general theory explaining the evolution of developmental bias will not only affect the rate or path toward an adaptive bias and its consequences—in particular, a theory that en- peak, but when adaptive landscapes are multipeaked, it can compasses nonlinear correlations, modularity, and other also change which peak is reached (Melo et al. 2016; Kounios forms of functional integration that have potential to facili- et al. 2017). Second, developmental systems can make phe- tate adaptation—remains to be articulated. As a conse- notypes develop in a correlated fashion even without reduc- quence, examples of bias may often be perceived as special ing or increasing variability in any individual trait (Pavlicev cases, idiosyncratic to specific taxa and thus interesting but of et al. 2011). Such correlations, responsible for the functional limited value for our fundamental understanding of the evo- integration of complex phenotypes, have the potential to lution of adaptation and diversification (Charlesworth et al. channel phenotypic variability toward directions of high fit- 1982; Maynard-Smith et al. 1985; Futuyma 2017). ness (Watson and Szathmary 2016; Figure 2C). We suggest that this conclusion is at best, premature, and A well-known example is the phenotypic integration of almost certainly mistaken. In what follows, we explain how vertebrate limbs (Hall 2015). Left and right hind limbs share the study of regulatory networks is beginning to reveal the developmental pathways, and mutations in a gene regulating evolutionary logic of developmental bias, including facilitated bone growth will therefore usually affect both limbs, making variation, and their profound consequences for understanding them grow equally.In the course of growth, bones themselves what determines the rate and direction of the evolutionary help instruct the development of muscles, tendons, and their process. This work implies that bias is likely to be the default respective attachment sites, ensuring that mutations in genes condition, and that consideration of bias will be highly in- that only directly affect skeletal growth nevertheless result in structive in evolutionary analyses. functional, well-integrated limbs. Further, bone growth re- sponds to mechanical pressure, which helps to accommodate Evolution of Developmental Bias: A Regulatory Network both genetic and environmental perturbations in ways that Perspective maintain functional integration within and between limbs. By preventing the expression of variants with longer limbs The historical treatment of bias as solely constraint and the on one side of the body, or variants that have mismatches associated focus on physical or material limits to biological between bones, muscles, and tendons, regulation of limb de- form (“universal constraints,” Maynard-Smith et al. 1985) velopment promotes variants in directions likely to be func- have distracted attention from how developmental biases tional, even under evolutionarily novel conditions (e.g., can evolve through natural selection. That mutational bias Standen et al. 2014). This is not a special case; the depen- can influence molecular evolution (e.g., Yampolsky and dencies between different components of development have Stoltzfus 2001; Nei 2013; Stoltzfus and McCandlish 2017), the potential to capture and channel random mutational var- and that selection can favor mechanisms that influence the iation toward nonrandom, functional, integrated, pheno- rate at which heritable variation arises (e.g., Charlesworth types. This, in turn, can make adaptive variants more easily 1976; Feldman and Liberman 1986; Day and Bonduriansky accessible to selection, and reduces the number of regulatory 2011; Geoghegan and Spencer 2012), are both well-established changes needed to convert developmental variation to evo- principles. However, understanding developmental bias lutionary, adaptive changes in form and function (West- requires attention not only to the frequency at which muta- Eberhard 2003; Kirschner and Gerhart 2005; Gerhart and tions arise, but also to the phenotypic properties of those Kirschner 2007). This line of reasoning has been interpreted variants. Most of the well-established tools of the evolution- by some (e.g., Félix 2016) to imply that most mutations would ary biologist are not well-designed to deal with the evolution produce functional phenotypes—a notion at odds with empir- of development, and shed limited light on how trait correla- ical observations. However, facilitated variation (Gerhart and tions originate and evolve (Rice 2004, 2008; Watson et al. Kirschner 2007) makes no such claim, but merely posits that 2016). Although developmental constraint has a long re- the interdependencies of developmental processes increase search tradition in evolutionary [re- the probability of directing the effect of mutations toward views in Arnold (1992), Cheverud (1996), Hansen and functional phenotypes able to fuel adaptive response to selec- Houle (2008); Box 2], the reliance of quantitative genetics tion more rapidly than would otherwise be the case. Facilitated on linear statistical correlations means that it struggles to variation is entirely consistent with the empirical observation adequately represent variation containing gaps and some that most genetic mutations are either neutral or deleterious, other nonlinear interactions otherwise common in develop- and that changes in amino acids most commonly disrupt the ment [Watson et al. 2014; but see Morrissey (2015)]. Many function of proteins. insights into the evolutionary causes and consequences of That the mechanisms of development facilitate rather than developmental bias therefore come from the representation merely constrain functional integration raises the possibility of phenotypic distributions using mechanistic models, such as

952 T. Uller et al. Figure 2 Developmental bias can both constrain and facilitate adaptive evolu- tion. (A–D) Adaptive landscapes with a ridge and a positive slope toward the top-right corner. The shading represents the distribution of evolutionarily relevant phenotypic variation introduced into the population (e.g., by mutation), with darker regions representing higher frequencies of variants. (A) The default assumption in evolutionary theory is typically that the distribution of evolu- tionarily relevant phenotypic variation in- troduced into the population is unbiased. (B) Developmental bias will constrain adaptive evolution if it limits variability in the direction of selection. (C) Develop- mental bias will accelerate adaptive evo- lution if it biases variability in dimensions aligned with the direction of selection. (D) Recent theory and empirical research described in this paper further suggests that developmental bias can itself evolve both to orient with the adaptive land- scape and to increase phenotypic vari- ability in the direction favored by past natural selection (dashed arrow repre- sents changes in phenotypic distribution over time as the population evolves).

regulatory networks, that can capture nonlinear relationships Britten and Davidson 1969; Davidson 2006) (Figure 3A). and multimodal distributions. Below, we explain how the Theinputtoanodemaybethepresenceoramountofa evolution of these networks can make nondirected genetic that regulates gene expression and the change or novel environments bias phenotypic variation to- output of a node the level of gene product. Interactions be- ward functional solutions. tween nodes can be described using linear or nonlinear functions. The of the network is the profile of gene expression of one or more nodes, which may represent The Phenotype as a Regulatory System the macroscopic phenotype of interest, such as morphology Biological processes, such as gene expression, metabolism, or physiology. However, networks are not restricted to gene and signaling cascades, lend themselves well to network- interactions. More explicit developmental models describe based models (Kauffman 1969; Alon 2006), and a major interactions at different levels of biological organization, role for regulatory changes in evolution is empirically well- such as cells and tissues (Oster and Alberch 1982; Atchley supported (Prud’homme et al. 2007; Wray 2007; Wittkopp and Hall 1991; Salazar-Ciudad et al. 2003; Newman and and Kalay 2012). Representation of the phenotype in terms of Müller 2005), whose dynamical changes feedback on tran- a network of interacting components with dynamical proper- scriptional regulation in multilayered models (von Dassow ties indeed has a long history,including seminal contributions et al. 2000; Salazar-Ciudad and Jernvall 2010). One important by Muller (1922), Schmalhausen (1949), Waddington (1957), feature of biological networks is that nodes (e.g., expression of and Kauffman (1969). Conceptualizing phenotypes as the out- genes) can be free to vary in their activity independently of put of regulatory networks remains prevalent among contem- each other, or can be connected by regulatory linkage. This porary developmental biologists (e.g., Wilkins 2005, 2007; allows networks to represent modularity (i.e., the extent to Davidson 2006; Peter and Davidson 2015). which different characters are developmentally integrated; In computational analyses, regulatory networks are often Schlosser 2002), which affects the statistical correlations of represented by genes as nodes and the suppression or acti- characters within a population (Cheverud 1996; Melo et al. vation of other genes as edges (gene regulatory networks; 2016).

Developmental Bias and Evolution 953 Box 2 Developmental bias in evolutionary quantitative genetics Quantitative genetics is a statistical approach to modeling phenotypic evolution. Its canonical equation is the multivariate breeder’s equation, Dz=Gb (Falconer and Mackay 1996; Lynch 1998). This equation describes evolutionary change in a suite of traits, described as a vector of differences in trait means, Dz, as the product of a vector of selection gradients, b, and a matrix, G, whose entries are the additive genetic variances and covariances of the traits. Correlational selection has a tendency to ensure that traits that are selected together are inherited together (Lande and Arnold 1983). The coinher- itance of traits at the population level is specified by their genetic covariance. As selection removes variants with low fitness, the genetic covariation in large populations will tend to be proportional to the patterns of mutational variance at pleiotropic loci and the strength of multivariate selection (Lande 1980). The genetic variance–covariance (i.e., G)isan estimate of the biasing effect on evolution of standing genetic variation (Arnold 1992). It has been suggested that the lead

eigenvector of G (gmax) predicts evolutionary trajectories because genetic variances and covariances constrain possible changes and hence the response to selection (e.g., Schluter 1996). However, G is of limited value for understanding developmental bias since the same pattern of genetic covariation can arise from a variety of distributions of and functional (e.g., Houle 1991; Gromko 1995). G describes currently existing variation but not the propensity to generate variation (variability). A more relevant entity for understanding bias is the distribution of mutational effects, which is how new mutations enter the population (Lande 1980; Cheverud 1984). The distribution of mutational effects, which is often called the M matrix, depends on patterns of pleiotropy and epistasis (e.g., Jones et al. 2007; Chebib and Guillaume 2017). Although most quantitative genetic theory assumes that mutations have uniform effects on the phe- notype (e.g., Lande 1980), both pleiotropy and epistasis are potentially evolvable features. How the elements of M evolve can be modeled if one assumes that they are underpinned by additive genetic variation (Jones et al. 2007; Pavlicev et al. 2011), or by modeling the evolution of pleiotropic loci connected by epistatic coefficients (Jones et al. 2014). A key finding of these models is that both stabilizing and directional correlational selection can result in patterns of pleiotropy and epistasis that align mutational effects with the direction of the fitness landscape [Pavlicev et al. 2011; but also see Hansen et al. (2006)]. Thus, new genetic variants may bias the phenotype in the direction favored by past selection (see Evolution of Facilitated Variation below; Figure 1, C and D).

In network models, “mutations” may represent changes in sometimes result in a qualitative shift in phenotype. Similarly, topology, such as the deletion or addition of a node or link, or even simple networks have multiple attractors that can a change of interaction from suppression to activation (Figure lead minor differences in the dynamic parameters to have 3A). Networks also have dynamical properties that describe large phenotypic consequences (Figure 3B). One such exam- the trajectory of the inputs and outputs as the system con- ple is the gap gene networks that regulate body segmentation verges (if at all) on one of possibly several steady states, or in the early fly embryo. By simulating the biological networks phenotypes (Figure 3B). The parameters that determine the in silico, Jaeger and colleagues have shown how changes in dynamical properties of networks include initial conditions, the concentration of maternally derived mRNA (i.e., initial the activating input to the nodes, and changes in the strength conditions) can cause different expression profiles across the of interactions (Jaeger and Monk 2014), all of which may body, sometimes resulting in different segmentation pheno- show heritable variation. types [Wotton et al. 2015; see also Clark (2017)]. Such analy- Given that there is often ample standing genetic variation ses illustrate how developmental mechanisms bias phenotypes in natural populations, it might seem that any bias must be toward particular outcomes, including producing the same transient and of little bearing on evolution (Charlesworth and phenotypes under a range of different starting conditions. Lande 1982, 2017; Coyne 2006; Futuyma 2015). Yet the From a network perspective, phenotypic evolution is typ- amount of standing genetic variation may have little rele- ically represented by heritable changes in the topology or vance for the probability that particular phenotypes will be dynamical properties of regulatory networks in the popula- generated. Studies of simulated and real regulatory networks tion. An important observation is that different topologies are demonstrate that only a small part of the phenotypic space often functionally equivalent (Kauffman 1983; Borenstein can be reached by a given developmental system, while the and Krakauer 2008; Wagner 2011). Although the remainder is inaccessible [Kauffman 1983; Borenstein and to mutation may at first seem to limit the potential for evo- Krakauer 2008; see also Dingle et al. (2015)]. Even within lution, theory suggests that it in fact increases the capacity to accessible regions of phenotype space, changes in topology, evolve (Fontana and Schuster 1998; Ciliberti et al. 2007; initial conditions, or interaction strengths will not always, or Wagner 2011). To understand how, one needs to consider even commonly, produce a smooth transition in phenotype two features of the space of possible regulatory networks (Kauffman 1969, 1983; Alon 2006; Jaeger and Monk 2014). (below we refer to topologies as “,” but emphasize Networks that differ in only a single type of interaction between that this does not imply that regulatory networks are solely two nodes, such as replacing activation with suppression, can represented in terms of genes).

954 T. Uller et al. Figure 3 Regulatory networks, their topol- ogy, dynamics, and connectivity via muta- tion. (A) A regulatory network with nodes (e.g., genes) connected by regulatory inter- actions can be considered a . Mu- tation to the genotype is represented by a modification in the network topology, for example, by adding or removing regulatory interactions or by modifying the interactions from suppression to activation. Pointed ar- rows represent activation, while those end- ing in perpendicular lines show suppression. (B) A regulatory network has a phase space, here represented by the concentrations of molecules encoded by the two genes. The flow of the phase space (gray arrows) de- scribes what trajectory (black arrow) from the starting point (red circles) the system will take as it reaches equilibrium (“basin of at- traction”; black circles). Which of potentially several equilibrium states is reached can de- pend on external conditions, such as the concentration of the activating substance (red arrow and red circles). (C) Connecting regulatory networks (A) to other regulatory networks that differ in only one regulatory change results in large “network-of-networks” or genotype networks. Real genotype net- works are very large so two dimensions can only represent a small part of all possible genotypes and how they are connected by mutation. In this hypothetical example, each node represents a single regulatory network, with the color indicating its phenotype. Edges connect regulatory networks that are related by a single modification of their to- pologies, as represented in A. The properties of these genotype networks determine how likely it is that an alternative phenotype can be reached through mutation. The three shaded areas represent, from left to right, (i) a boundary region between two distinct phenotypes where some genotypes can pro- duce both phenotypes; (ii) a region of geno- type space where mutations (i.e., changes in topology) are phenotypically neutral; and (iii) a region where a change in topology can produce several distinct phenotypes that are not accessible from other parts of the network. (D) The regulatory network of this node in the genotype network. ‘A and B are based on representations in Jaeger and Crombach (2012) and Jaeger and Monk (2012) and C is based on representations in Wagner (2011).

The first is that genotypes that share the same phenotype The fact that vast areas of the space of possible regulatory are often topologically similar, forming a “network-of- networks are phenotypically equivalent, but have different networks” (or a “genotype network space”) where each ge- phenotypic neighbors, is important because a population that notype is connected to other genotypes by a single addition, harbors many genotypes (i.e., regulatory networks) with the deletion, or modification of one of the regulatory network same phenotypes can more easily find a new phenotype components (Fontana and Schuster 1998; Wagner 2011; Fig- through a single mutational step (Wagner 2011; Figure 3C). ure 3C). The functional equivalence of neighboring topolo- Since the number of neighboring genotypes with different gies allows populations to accumulate genotypic variation phenotypes increases with the absolute number of neighbors, through drift. The second is that regulatory networks that the capacity for a population to evolve new functional pheno- are topologically quite different can nevertheless share the types should be higher if there are many phenotypically neu- same phenotype (Wagner 2011). However, because they are tral neighbors than it would be if every change to a regulatory topologically different, the phenotypes of their neighbors network produced a different phenotype (Wagner 2011). (i.e., a network differing in only one regulatory change) Whether or not developmental bias, in the form of the may be radically different. phenotypic consequences of genetic perturbation, is expressed

Developmental Bias and Evolution 955 within a particular population depends on which neighboring not necessarily facilitate adaptation to environments that are regulatory networks can be reached through mutation. That, in structurally different (Kouvaris et al. 2017). To study the turn, depends on the population’s evolutionary history. Stabi- latter, Parter et al. (2008) modeled a combinatorial logic lizing selection will tend to push the population to regions of circuit and a secondary RNA structure using nonlinear inter- genotype space where changes in topology do not affect the actions that allow small changes in genotype to produce large phenotype (Ciliberti et al. 2007; Wagner 2011), which favors changes in phenotype [see also Kashtan and Alon (2005), regulatory networks with robust dynamical properties such as Kashtan et al. (2007)]. When evolved under conditions insensitivity to fluctuations in transcription factor concentra- where selection switched between two “goals” (i.e., target tions (Jaeger and Monk 2014). In contrast, directional or dis- phenotypes matched to particular environments) on the or- ruptive selection can push the population toward regions of der of tens of generations, the model was soon reliably able to genotype space where small changes in regulatory network evolve to match both goals. More importantly,the system was topology or its parameters are more likely to produce pheno- able to evolve adaptive phenotypes in environments that typic effects (Kashtan et al. 2007). Theoretical analyses sug- shared the same structural regularity but that had not been gest that such differences in genotype neighborhoods are previously encountered (including those requiring novel likely, and that they affect the likelihood that populations combinations of subgoals). Similar results are found in gene can find new high-fitness phenotypes through small changes regulatory network models evolving in fluctuating environ- in regulatory network topology (Psujek and Beer 2008; Payne ments (Crombach and Hogeweg 2008; Draghi and Wagner et al. 2014). For example, a comprehensive theoretical study of 2009; Watson et al. 2014; Kouvaris et al. 2017). three-gene circuit topologies that produce a stripe demon- At first sight, that regulatory networks evolve features that strated that, of the thousands of possible circuits, each rely then allow them to adapt quickly to conditions that they have on one of only six mechanisms, which differ in their likelihood not previously encountered appears incompatible with the to reach new phenotypes through mutation (Jiménez et al. myopic vision of natural selection that rewards current and 2015). While further quantification of developmental bias not future function (Watson and Szathmary 2016; Kounios across biological systems is necessary to establish how impor- et al. 2017). To understand these results, it is necessary to tant bias is to real populations, these theoretical analyses illus- revisit some of the properties of networks described above. trate that bias is likely to be an inherent property of evolved First, remember that under a stabilizing selection process, regulatory systems. the population will evolve toward regulatory networks that have a large mutational distance to other phenotypes (i.e., toward the center of the genotype network with the particu- Evolution of Facilitated Variation lar phenotype, Wagner 2011; Figure 3C). The same logic Perhaps the most surprising finding from studies of the evo- implies that switching between two environments at a fre- lution of regulatory networks is that phenotypic variability can quency that enables populations to adapt but not to evolve be directed toward dimensions with high-fitness variance regulatory networks that are mutationally robust, will tend to even when mutations are randomly distributed. For example, push genotypes toward a space of possible regulatory net- allowing interaction between genes controlling two traits to works where the mutational distance is short between net- evolve under stabilizing or directional correlational selection works that are functional in environment one and networks causes the mutational effects (i.e., the M matrix) to become that are functional in environment two (Kashtan et al. 2007; biased toward phenotypes that are aligned with the fitness Wagner 2011; Figure 3C). The regulatory networks on the landscape (Jones et al. 2007, 2014; Pavlicev et al. 2011; boundary that are performing best have some nodes or edges Watson et al. 2014). The reason is that selection strengthens that have disproportionate effects on the phenotypic outcome the interactions between genes (i.e., epistatic effects) that (Parter et al. 2008). produce the desired correlation among characters, while it Second, evolution in a structurally complex environment reduces the strength of interactions among genes that pro- can favor regulatory networks that are modular (Lipson et al. duce undesired correlation. If trait correlations evolve more 2002; Clune et al. 2013; Kouvaris et al. 2017). In the simu- slowly than the quantitative traits themselves, the combina- lations performed by Parter et al. (2008), each goal was dis- tion of trait values under new mutations will be biased to- tinct but composed of different combinations of the same set ward those that have been favored in the past. The result of of subgoals. Switching between two goals (each consisting of this mutational effect is an increase in the standing popula- different combinations of the same subgoals) makes the net- tion genetic covariation between traits (i.e., the G matrix) in work evolve modularity (Kashtan et al. 2007; Crombach and the dimension(s) aligned with past selection. Hogeweg 2008; Clune et al. 2013; Watson et al. 2014). With With linear interactions between traits, such developmen- a modular network topology, mutations within modules can tal interactions can facilitate the production of functional have a relatively large but specific phenotypic effect, which phenotypes in more extreme environments because of the enhances the possibility to acquire novel functions while re- correlational selection in the environments to which the ducing the pleiotropic effect of mutation on other modules. system has adapted (e.g., Draghi and Whitlock 2012). This makes it possible for regulatory networks to use their This can accelerate adaptive evolution, although it would modular structure to evolve new topologies that perform well

956 T. Uller et al. in environments that are novel, but that retain underlying to explain biological features that are difficult to account for features of past environments (Clune et al. 2013; Kouvaris assuming that selection acts on unbiased variation. Such et al. 2017). features include the rapid adaptation of complex pheno- These properties of regulatory networks evolving under types, why some lineages continue to diversify while others natural selection suggests that evolution exploits the under- do not, and why some features evolve repeatedly, often- lying structural regularity of the environment to produce times using the same developmental pathways, whereas developmental systems that retain a bias toward phenotypes others are one-offs. Below, we brieflydiscusskeycompo- evolved in the past (Lipson et al. 2002; Watson et al. 2014). As nents of the relationship between developmental bias and a result, evolving systems can exhibit bias toward phenotypes evolution. that are fit even in environments that have not been previ- Developmental bias can influence taxonomic and ously encountered, exploiting their modular structure (Parter phenotypic diversity et al. 2008; Watson and Szathmary 2016; Kouvaris et al. 2017). If future environments are structurally similar to Evolutionary change in regulatory interactions may help to those of the past, bias should facilitate adaptive evolution, explain some puzzling observations with respect to the accu- whereas it should limit adaptation in structurally different mulation of phenotypic diversity through time (McShea 1994; environments. Erwin 2017; Jablonski 2017). Low-dimensional regulatory Although most models focus on genetic change to regula- networks have been found to produce higher disparity among tory networks, environmental perturbation may also be an common phenotypes than high-dimensional networks important source of developmental bias, not least because (Borenstein and Krakauer 2008), suggesting that diversifica- organisms may be more likely to have evolved adaptive tion rate will be highest early in evolutionary time when responses to environmental than genetic variation [reviewed regulatory networks are small. Such models predict that line- in West-Eberhard (2003), Pfennig et al. (2010), Moczek et al. ages will become increasingly clumped as evolution prog- (2011), Levis and Pfennig (2016), Schneider and Meyer resses, with the greatest divergences appearing early as (2017)]. Even if environmentally induced phenotypes are higher-level taxonomic grades (Salazar-Ciudad and Jernvall not heritable, plasticity has the potential to facilitate adapta- 2005; Borenstein and Krakauer 2008). These predictions are tion by increasing the recurrence and fitness of functional consistent with the early bursts of radiation seen across sev- variants, which tends to increase their likelihood of being se- eral metazoan taxa, including tetrapods and arthropods lected and reduce the amount of genetic change needed to (Davidson and Erwin 2006; Hughes et al. 2013; more com- convert them into locally adapted phenotypes (Waddington plex patterns have also been described, e.g., Wright 2017). 1957; West-Eberhard 2003; Gerhart and Kirschner 2007). Preliminary studies suggest a similar pattern for some plants Compared to regulatory networks represented by genes alone, (Oyston et al. 2016). It has even been suggested that the networks with environmental dependencies have been dem- rapid evolutionary diversification of body plans during the onstrated to evolve greater modularity, increased mutational Cambrian explosion were caused by the evolution of par- distance to phenotypically disparate networks, and mutational ticular gene regulatory networks (“kernels”; Davidson and variance that is exaggerated in the direction of past selection Erwin 2006). (Espinosa-Soto et al. 2011; Fierst 2011; Wagner 2011; Draghi Within lineages, the evolution of novelties, such as shells, and Whitlock 2012; van Gestel and Weissing 2016), these limbs, photic organs, feathers, wing patterns, or horns, is being features associated with enhanced evolvability. The rich associated with rewiring existing developmental building literature on the effects of learning on evolution provides fur- blocks and processes into new regulatory networks. This ther insights into how plasticity contributes to bias and evolv- predictsthat,oncetheyappear,diversification of novelties ability (Box 3). should proceed rapidly at first,andslowdownastheir regulation becomes developmentally entrenched. Consis- tent with this prediction, the shape of bird bills diverged Detecting Signatures of Developmental Bias in Phenotypic rapidly during the early radiation of modern birds, and Evolution subsequent evolution of bill shapes within major bird line- Natural selection and developmental bias (or constraint) ages has been filling up only limited parts of morphospace have often been pitted against each other as alternative (Cooney et al. 2017). Mathematical analyses of the morpho- explanations for phenotypic variation, but such a juxtapo- space of bird bills and experimental manipulation of bill sition is misleading. The recognition of developmental bias growth indeed demonstrate that much of the observed di- does not change the status of natural selection, which versityinshapecanbeexplainedbychangesinonlyafew remains the process by which some variants are retained parameters that describe regulatory interactions among key and others are removed as a result of fitness differences genes (Campas et al. 2010; Mallarino et al. 2011; Fritz et al. between individuals. However, the theory reviewed above 2014), suggesting that much of the remaining parts of mor- demonstrates that the phenotypic variation exposed to se- phospace is empty as a result of how bill development is lection will reflect the lineage’s evolutionary history. The regulated. The evolutionary fixationofgeneregulatorynet- explanatory value of developmental bias is that it can help works has been applied more generally to explain why

Developmental Bias and Evolution 957 Box 3 Learning, developmental bias, and evolvability As a form of adaptive plasticity that allows organisms to shift their phenotype toward the optimum, learning is inherently a source of developmental bias. Learned behavior is often the result of an exploratory search conducted over multiple trials, and this search is expanded to encompass the experiences of multiple individuals where animals learn socially. Extensive theory has demonstrated that learning has an advantageous effect on adaptation in changing environments, allowing individuals to acclimate to changes that cannot be tracked by selection of genes (Cavalli-Sforza and Feldman 1981; Boyd and Richerson 1985; Todd 1991). More contentious are the benefits of learning in stationary or slowly changing envi- ronments. Hinton and Nowlan (1987) suggested that learning could accelerate evolution in a static environment by helping genotypes to locate otherwise difficult-to-find fitness peaks. However, learning is also known to weaken selection by reducing phenotypic differences between genotypes (Anderson 1995; Ancel 2000; Frank 2011). The conflicting findings follow from different assumptions about the structure of fitness landscapes (Borenstein et al. 2006; Paenke et al. 2007; Frank 2011). The emerging consensus from theoretical analyses is that individual learning typically slows evolution in static unimodal fitness landscapes, but usually accelerates evolution in dynamic or static multimodal fitness landscapes. In the latter, the existence of multiple optima usually slows down the evolutionary process as populations become trapped on suboptimal fitness peaks. By smoothing the landscape, learning increases the likelihood of a directly increasing path of fitness to the global optimum (Borenstein et al. 2006; Mills and Watson 2006; Frank 2011). These findings parallel analyses using gene regulatory networks that, in contrast to more traditional reaction–norm modeling frameworks, also found that adaptive plasticity can reduce the likelihood of getting stuck on local fitness peaks (van Gestel and Weissing 2016; Kounios et al. 2017). More generally, diverse forms of phenotypic plasticity operate in a functionally equivalent manner to learning, by relying on a combination of exploratory and selective processes (e.g., adaptive immune system, vascular system, nervous system) (Gerhart and Kirschner 2007; Snell-Rood 2012). Such processes are thought to allow organisms to respond to evolutionarily novel environmental challenges in a manner that generates phenotypic variation aligned with functional demands. This raises the possibility that the theoretical findings concerning the de- velopmental bias arising from learning may generalize to a broader class of adaptive plasticity. particular features of organisms are conserved and how mathematical model derived on the basis of knowledge of developmental regulation channels phenotypic variation the mechanisms of tooth production in mice could be used (Wagner 2014). to predict the relative sizes of teeth in a sample of 29 other rodent species. Herbivores tended to have more equal sized Developmental bias can impose directionality teeth and carnivores less equal, but all species were positioned on evolution along the same dimension of morphological space. Such stud- Testing the prediction that divergence between lineages is ies raise the possibility that natural selection may only be able fi shaped by the variational properties of development is chal- to move species along highly speci c pathways created by the lenging, and would ideally be substantiated by a detailed mechanisms of development. knowledge of . Here, we highlight a In the absence of models that can predict patterns of small number of studies whose results are consistent with the variability, empiricists are often limited to comparing pheno- theoretical prediction that the direction of phenotypic change types within populations or species with the pattern of phe- fi over evolutionary time will be concordant with, and hence notypic diversi cation across species (Klingenberg 2014; sometimes can be predicted by developmental bias. Goswami et al. 2015). Although this risks confounding vari- Tooth morphology in mammals has both awell-understood ation and variability (Box 1), phenotypic covariance in developmental biology and a detailed record of evolutionary morphological characters in extant vertebrates has been diversification. In the computational model of Salazar-Ciudad demonstrated to be concordant with the patterns of his- and Jernvall (2010), tens of parameters describing known torical diversification, including for example pharyngeal genetic and cellular interactions were modeled, with modifi- jaw morphology in cichlids (Muschick et al. 2011), beak cation of only one or a few of these accurately predicting evo- shape in raptors (Bright et al. 2016), skull morphology in lutionary diversification of teeth across several groups of toads (Simon et al. 2016), and body shape in sticklebacks mammals [Salazar-Ciudad and Jernvall 2010; Harjunmaa (Schluter 1996). et al. 2014; see also Kavanagh et al. (2007), Evans et al. More robust inference is possible through complementary (2016)]. These models not only support the view that the studies of the effects of genetic mutation. A particularly evolutionary diversity in tooth morphology among mammals impressive study used data on wing shape for over 50,000 is shaped by the mechanism by which teeth develop, but they fruit flies to study the relationship between the phenotypic also generate predictions for what developmental and ge- changes caused by mutation, standing genetic variation, and netic changes should accompany adaptive diversification of disparity among species (Houle et al. 2017). Despite the fact teeth. For instance, Kavanagh et al. (2007) showed that a that mutations occur much more frequently than necessary to

958 T. Uller et al. account for the phenotypic divergence between species, the pathways, and morphogenetic processes was used to arrive at phenotypic variants introduced by mutation within species functionally highly similar outcomes. Rather than reflecting parallel the phenotypic disparity between species. One inter- constraint, such cases are consistent with developmental sys- pretation of these results is that, as predicted by the regula- tems shaping evolutionary trajectories by generating oppor- tory network models described above, the propensity to vary tunities to evolve complex structures repeatedly, reliably and in response to mutation is coevolving with the phenotypic regardless of taxonomic context. At the same time, the num- divergence between species (Cheverud 2017). A similar ber of genetic changes needed to evolve a lineage-specific study of developmental variation in the nematode vulva also eye, heart, or appendage is significantly reduced compared found that differences between genera in the covariation to a scenario requiring the de novo evolution of genes for each among characters caused by mutation was concordant with structure. how vulva morphology have diversified (Todd and Miller Nevertheless, distinguishing between bias that constrained 1991; Dichtel et al. 2001; Kiontke et al. 2007; Braendle evolution and bias that facilitated adaptation is challenging. et al. 2010). The wealth of information on the developmental Particularly promising examples illustrating the existence and biology of the nematode vulva and the Drosophila wing make significance of the latter are where plastic responses that help them outstanding cases for mechanistic models that can in- organisms cope in stressful environments become genetically vestigate whether the patterns of developmental bias are accommodated (West-Eberhard 2003). Evolution via genetic consistent with the mechanisms of development (Félix and accommodation of plastic responses has been demonstrated Barkoulas 2012; Matamoro-Vidal et al. 2015). experimentally (e.g., Suzuki and Nijhout 2006), and a biasing Another detailed example comes from studies of the size, effect of phenotypic plasticity within populations or species is position, and color of eye spots in Mycalesine butterflies. A known to mirror patterns of evolutionary diversification in a combination of artificial selection and quantification of the diversity of taxa. In both cichlids and sticklebacks, the mor- variation observed within and among species have revealed phology of the feeding apparatus that develop when indi- that characters that show little evidence for bias within species viduals are reared on a food source to which they are not (i.e., size of different eye spots, which respond readily to adapted resembles the morphology observed in species selection; Beldade et al. 2002) exhibit a diversity across spe- adapted to the same food (Wund et al. 2008; Muschick cies that fills up a large portion of morphospace (Brakefield et al. 2011). This suggests that evolution has capitalized on and Roskam 2006). Conversely, characters that show clear the effects of physical stress whose functionality was ensured evidence for bias within species (i.e., eye spot color, which by channeling cellular and genetic regulatory networks in shows much more limited variability and fails to respond morphogenesis. Environmentally induced bias has also been to antagonistic selection; Allen et al. 2008) show a corre- suggested as a contributor to the evolution of carotenoid sponding limited diversity across the Mycalesine butterflies coloration in birds (Badyaev et al. 2017), pigmentation in (Brakefield 2010). The differences in evolvability between water fleas (Scoville and Pfrender 2010), morphology and eyespotsizeandcolorationappeartoreflect variability physiology in carnivorous toads (Gomez-Mestre and Buchholz of the underlying developmental mechanisms. Importantly, 2006; Kulkarni et al. 2017), morphological and behavioral while the majority of species fit with the trends predicted traits in Onthophagus dung beetles (Casasa and Moczek based on knowledge of developmental mechanism, certain 2018), and sexual size dimorphism in the house finch exceptional species were found with eye spots that did not (Badyaev 2005). match expectations, a situation that also applies to the study of mammalian teeth. Such findings suggest that organisms Challenges and Opportunities for Future Studies may most often fall along a developmentally favored evolu- tionary trajectory but that developmental bias need not im- Longstanding controversy over the roles of developmental pose constraints that are impossible to break (Kavanagh et al. constraints and bias in evolution reflects both conceptual and 2007; Brakefield 2010). methodological challenges (e.g., Maynard-Smith et al. 1985; The potential macroevolutionary significance of develop- Amundson 2005; Salazar-Ciudad 2008). The representation mental bias is further exemplified by hundreds of examples of of phenotypes in terms of regulatory networks resolves some repeated co-option and recruitment of the same developmen- of the contention as it helps to explain how evolution can give tal pathways into the building of analogous structures and rise to developmental bias even when bias itself is not a target organs in otherwise unrelated organisms [reviewed in Shubin of selection. However, the evolutionary consequences of bias et al. (2009) and Held (2017)]. Some of the most spectacular are important even if developmental bias has been favored by cases include the independent evolution of eyes across phyla selection. In both cases, the propensity to vary is expected to (Mercader et al. 1999; Kozmik 2005; Kozmik et al. 2008), the be coevolving with the phenotypes themselves. Thus, the evolution or contractile hearts in vertebrates and inverte- contributions of natural selection and developmental bias brates (Olson 2006; Xavier-Neto et al. 2007), or the forma- to adaptation and diversification are not easily quantified tion of outgrowths from insect legs to echinoderm tube feet or disentangled. or ascidian siphons (Panganiban et al. 1997; Mercader et al. One useful approach would be to identify conditions under 1999); in each set of cases the same set of preexisting genes, which evolution with bias should proceed differentially from

Developmental Bias and Evolution 959 Table 1 Evolutionary questions that the study of developmental bias helps to answer Question Answer with key reference

Why is the influence of genetic and environmental The feedback, modular structure, and nonlinear interactions of regulatory networks allow change on phenotypes not uniform? developmental systems to exhibit both robustness (i.e., no or small phenotypic change even under large perturbation) and innovation (i.e., large yet functionally integrated phenotypic change even under small perturbation) (Wagner 2011). How can regulatory networks facilitate the expression As regulatory interactions evolve, they discover underlying structural regularities of the of functional phenotypes when populations are environments to which they become adapted, including through modular structure, making exposed to novel environments? it possible to reach new adaptive combinations of characters through a small number of mutations (Watson and Szathmary 2016). Why did a great deal of morphological variation Simple, low-dimensional ancestral regulatory networks will tend to produce greater disparity among evolve early in the history of multicellular life? the set of common phenotypes than derived high-dimensional networks because ancestral genotypes are less constrained by regulatory epistasis (Borenstein and Krakauer 2008). Why do phenotypes occupy only a small region of Chance and the adaptive demands of natural selection combine with regulatory epistasis in possible phenotype space? evolving networks to leave only a fraction of possible phenotypes reachable (Wagner 2011). How can developmental processes influence the Evolution of regulatory networks illustrate that the phenotypic variation available for natural direction of phenotypic evolution? selection will typically be biased, sometimes in a functional manner, even when mutations are randomly distributed (Watson and Szathmary 2016). How does developmental bias contribute to Developmental bias increases the recurrence and fitness of new phenotypes, thereby evolvability? reducing the amount of genetic change needed to convert them into adaptive phenotypes (Watson and Szathmary 2016). Developmental bias may thus increase evolvability by making it more likely that adaptive phenotypes arise. How does developmental bias shape macro- Analyses of regulatory networks reveals that stabilizing selection will push evolving populations evolutionary patterns? to regions of genotype space where changes in topology do not affect the phenotype (generating stasis), while disruptive selection shifts populations to regions in which rapid change can ensue (Wagner 2011). This table provides only brief summary statements. Readers are referred to the main text for full explanations. evolution in the absence of bias, or a different bias. Surpris- likely to be produced. Comparative phylogenetic tools could ingly few models are designed to generate such predictions be usefully combined with this approach to ascertain explicitly (Kovaka 2017). The rapid increase in the number of whether bias was present in ancestral lineages. Such com- studies that provide compelling empirical evidence that pat- parative statistical methods can validate, or be validated by terns of phenotypic diversification can be concordant with experimental approaches to investigating bias that allow developmental bias make the development of such theory retrieval of ancestral character states (Harjunmaa et al. all the more relevant. 2014). Studies that combine experimental work on devel- Although the patterns of temporal and extant interspecies opmental mechanisms with phylogenetic reconstructions diversity can be consistent with a contribution of develop- and surveys of occupancy of morphospace may be able to mental bias, alternative explanations must be considered. For reveal if developmental bias contributes to the patterns of example, selection could have produced temporal patterns of convergence often apparent in parallel adaptive radiations, diversification from unbiased variation if the appearance of as in the anoles of Caribbean and African lake cichlids ecological niches was nonuniform and the filling of niches by (Losos 2011; Brawand et al. 2014). divergent lineages limited the opportunity for disparity within lineages (Pie and Weitz 2005). However, the fact that ecol- Conclusion ogy obviously affects rates and patterns of diversification (Schluter 2000; Rabosky 2009; Losos 2010) does not render In a seminal contribution to the study of developmental bias, the evolutionary effect of developmental bias unimportant or Pere Alberch (1989, p. 48) wrote: “The reason why develop- unresolvable (Brakefield 2006). If the evolutionary effect of ment has not been integrated into the existing corpus of evo- developmental bias is persistent, phenotypic diversification lutionary theory is not a technical one (the ‘we do not know into new ecological opportunities (e.g., through colonization enough about development’ type of argument) but a philo- of new environments) should remain channeled along evo- sophical one.” Our review of the above literatures suggests lutionary trajectories that are developmentally favored, be that, while the technical challenges are real, there is much disproportionally filled by organisms with the appropriate merit to Alberch’s analysis. Habits of thought—such as that variability, or even left unexploited. Comparisons of patterns bias can be understood as constraint; that bias results primar- of divergence over time between lineages or experimental ily from physical or material limitations on form; that it is populations that differ in their variability in ecologically rare, exceptional, or onerous; and that it provides an al- relevant characters could therefore provide important in- ternative explanation to selection—have all hindered the in- sights. As demonstrated by analyses of mammalian teeth, a tegration of development and evolution. The mounting knowledge of developmental mechanisms allows apriori evidence that phenotypic evolution commonly involves predictions to be made about the form of the variants most changes in the interactions among genes, cellular components,

960 T. Uller et al. cells, tissues, as well as organisms and their environments Allen, C. E., P. Beldade, B. J. Zwaan, and P. M. Brakefield, provides a strong impetus for evolutionary theory to address 2008 Differences in the selection response of serially repeated and incorporate how developmental systems acquire and con- color pattern characters: standing variation, development, and evolution. BMC Evol. Biol. 8: 94. trol the capacity to vary, and how variability affects the rate Alon, U., 2006 An Introduction to Systems Biology: Design Princi- and direction of evolution. The explanatory potential of de- ples of Biological Circuits. Chapman & Hall/CRC, Boca Raton, FL. velopmental processes for evolutionary biology remains re- Amundson, R., 2005 The Changing Role of the Embryo in Evolu- gardless of whether a given bias has itself been shaped by tionary Thought. Cambridge University Press, New York. natural selection or emerges through other processes. Thus, https://doi.org/10.1017/CBO9781139164856 fi Ancel, L., 2000 Undermining the Baldwin expediting effect: does it is not suf cient to accommodate developmental bias into phenotypic plasticity accelerate evolution? Theor. Pop. Biol. 58: evolutionary theory merely as a constraint on evolutionary 207–319. adaptation. Knowledge of the mechanisms that produce select- Anderson, R. W., 1995 Learning and evolution - a quantitative able phenotypes can afford both a more detailed understand- genetics approach. J. Theor. Biol. 175: 89–101. https://doi.org/ ing of patterns of variation found in nature, and of the 10.1006/jtbi.1995.0123 Arnold, S. J., 1992 Constraints on phenotypic evolution. Am. Nat. evolutionary dynamics of populations. With increasing recog- 140: S85–S107. https://doi.org/10.1086/285398 nition that the evolutionary process itself evolves (Watson and Arnold, S. J., M. E. Pfrender, and A. G. Jones, 2001 The adaptive Szathmary 2016; Watson et al. 2016), a consideration of bias landscape as a conceptual bridge between micro- and macro- promises the resolution of longstanding puzzles within evolu- evolution. Genetica 112–113: 9–32. https://doi.org/10.1023/ tionary biology (Table 1). A:1013373907708 Arthur, W., 2004 The effect of development on the direction of evolution: toward a twenty-first century consensus. Evol. Dev. 6: 282–288. https://doi.org/10.1111/j.1525-142X.2004.04033.x Atchley, W. R., and B. K. Hall, 1991 A model for development and Footnotes evolution of complex morphological structures. Biol. Rev. Camb. – 1Our use of the term “development” is as a synonym to ontog- Philos. Soc. 66: 101 157. https://doi.org/10.1111/j.1469-185X. 1991.tb01138.x eny, and is not intended to imply that adult forms are static; Badyaev, A. V., 2005 Maternal inheritance and rapid evolution of adult physiology and behavior can also generate bias. In the sexual size dimorphism: passive effects or active strategies? Am. context of this article, “development” is best seen as a shorthand Nat. 166: S17–S30. https://doi.org/10.1086/444601 for changes that occur to individuals during their life time. Badyaev, A. V., 2009 Evolutionary significance of phenotypic ac- 2We recognize two uses of the term “bias” in the literature, commodation in novel environments: an empirical test of the “ ” . Philos. Trans. R. Soc. Lond. B Biol. Sci. 364: which can be summarized as bias as process (i.e., the devel- 1125–1141. https://doi.org/10.1098/rstb.2008.0285 opmental processes that result in biased distributions of pheno- Badyaev,A.V.,A.L.Potticary,andE.S.Morrison,2017 Most types) and “bias as product” (i.e., the phenotypes themselves). colorful example of ? Exploring the While this duality of usage is potentially a source of confusion, in evolutionary destiny of recurrent phenotypic accommo- – practice the intended meaning is usually clear, given the context. dation. Am. Nat. 190: 266 280. https://doi.org/10.1086/ 692327 Beldade, P., K. Koops, and P. M. Brakefield, 2002 Developmental fl Acknowledgments constraints vs. exibility in morphological evolution. Nature 416: 844–847. https://doi.org/10.1038/416844a We are grateful to the members of our research groups for Borenstein, E., and D. C. Krakauer, 2008 An end to endless forms: discussions and to Nathalie Feiner, Miguel Brun-Usan, and epistasis, phenotype distribution bias, and nonuniform evolu- tion. PLoS Comput. Biol. 4: e1000202. https://doi.org/10.1371/ Alfredo Rago for constructive comments on a draft version of journal.pcbi.1000202 this manuscript, Adam Wilkins and two anonymous reviewers Borenstein, E., I. Meilijson, and E. Ruppin, 2006 The effect of for helpful comments on the submitted version of this phenotypic plasticity on evolution in multipeaked fitness land- manuscript, and Katrina Falkenberg for help with production scapes. J. Evol. Biol. 19: 1555–1570. https://doi.org/10.1111/ of the figures. This manuscript is the product of a collaboration j.1420-9101.2006.01125.x “ Boyd, R., and P. J. Richerson, 1985 Culture and the Evolutionary funded by the John Templeton Foundation ( Putting the ex- Process. Chicago University Press, Chicago. tended evolutionary synthesis to the test”,grantno.60501). Braendle, C., C. F. Baer, and M. A. Felix, 2010 Bias and evolution T.U. is supported by the Knut and Alice Wallenberg Founda- of the mutationally accessible phenotypic space in a develop- tions. A.P.M. is supported by the National Science Foundation. mental system. PLoS Genet. 6: e1000877. https://doi.org/10. 1371/journal.pgen.1000877 Brakefield, P. M., 2006 Evo-devo and constraints on selection. Trends Ecol. Evol. 21: 362–368. https://doi.org/10.1016/j. Literature Cited tree.2006.05.001 Brakefield, P. M., 2008 Prospects of evo-devo for linking pattern and Alberch, P., 1989 The logic of monsters: evidence for internal process in the evolution of morphospace in Evolving Pathways: Key constraint in development and evolution. Geobios Mem. Spec. Themes in Evolutionary Developmental Biology,editedbyA.Minelli, 22: 21–57. https://doi.org/10.1016/S0016-6995(89)80006-3 and G. Fusco. Cambridge University Press, Cambridge, UK. Alberch, P., and E. A. Gale, 1985 A developmental analysis of an Brakefield, P. M., 2010 Radiations of mycalesine butterflies and evolutionary trend: digital reduction in amphibians. Evolution opening up their exploration of morphospace. Am. Nat. 176: 39: 8–23. https://doi.org/10.1111/j.1558-5646.1985.tb04076.x S77–S87. https://doi.org/10.1086/657059

Developmental Bias and Evolution 961 Brakefield, P. M., and J. C. Roskam, 2006 Exploring evolutionary Clune, J., J. B. Mouret, and H. Lipson, 2013 The evolutionary constraints is a task for an integrative evolutionary biology. Am. origins of modularity. Proc. Biol. Sci. 280: 20122863. Nat. 168: S4–S13. https://doi.org/10.1086/509049 Cooney, C. R., J. A. Bright, E. J. R. Capp, A. M. Chira, E. C. Hughes Brawand, D., C. E. Wagner, Y. I. Li, M. Malinsky, I. Keller et al., et al., 2017 Mega-evolutionary dynamics of the adaptive radi- 2014 The genomic substrate for adaptive radiation in African ation of birds. Nature 542: 344–347. https://doi.org/10.1038/ cichlid fish. Nature 513: 375–381. https://doi.org/10.1038/ nature21074 nature13726 Coyne, J. A., 2006 Comment on “Gene regulatory networks and Bright, J. A., J. Marugan-Lobon, S. N. Cobbe, and E. J. Rayfield, the evolution of animal body plans”. Science 313: 761. 2016 The shapes of bird beaks are highly controlled by non- Crombach, A., and P. Hogeweg, 2008 Evolution of evolvability in dietary factors. Proc. Natl. Acad. Sci. USA 113: 5352–5357. gene regulatory networks. PLoS Comput. Biol. 4: e1000112. https://doi.org/10.1073/pnas.1602683113 https://doi.org/10.1371/journal.pcbi.1000112 Britten, R. J., and E. H. Davidson, 1969 Gene regulation for Darwin, C., 1859 On the Origin of Species. John Murray, London. higher cells - a theory. Science 165: 349–357. https://doi.org/ Davidson, E. H., 2006 The Regulatory Genome. Academic Press, 10.1126/science.165.3891.349 San Diego. Busey, H. A., E. E. Zattara, and A. P. Moczek, 2016 Conservation, Davidson, E. H., and D. H. Erwin, 2006 Gene regulatory networks innovation, and bias: embryonic segment boundaries position and the evolution of animal body plans. Science 311: 796–800. posterior, but not anterior, head horns in adult beetles. J. Exp. https://doi.org/10.1126/science.1113832 Zoolog. B Mol. Dev. Evol. 326: 271–279. https://doi.org/10. Day, T., and R. Bonduriansky, 2011 A unified approach to the 1002/jez.b.22682 evolutionary consequences of genetic and nongenetic inheritance. Camara, M. D., and M. Pigliucci, 1999 Mutational contributions Am. Nat. 178: E18–E36. https://doi.org/10.1086/660911 to genetic variance-covariance matrices: an experimental ap- Dichtel,M.L.,S.Louvet-Vallee,M.E.Viney,M.A.Felix,andP.W. proach using induced mutations in Arabidopsis thaliana. Evolu- Sternberg, 2001 Control of vulval cell division number in the nem- tion 53: 1692–1703. https://doi.org/10.1111/j.1558-5646. atode Oscheius/Dolichorhabditis sp CEW1. Genetics 157: 183–197. 1999.tb04554.x Dichtel-Danjoy, M. L., and M. A. Felix, 2004 Phenotypic neighbor- Camara, M. D., C. A. Ancell, and M. Pigliucci, 2000 Induced mu- hood and micro-evolvability. Trends Genet. 20: 268–276. tations: a novel tool to study phenotypic integration and evolu- https://doi.org/10.1016/j.tig.2004.03.010 tionary constraints in Arabidopsis thaliana. Evol. Ecol. Res. 2: Dingle, K., S. Schaper, and A. A. Louis, 2015 The structure of the 1009–1029. genotype-phenotype map strongly constrains the evolution of Campas, O., R. Mallarino, A. Herrel, A. Abzhanov, and M. P. Brenner, non-coding RNA. Interface Focus 5: 20150053. https://doi. 2010 Scaling and shear transformations capture beak shape org/10.1098/rsfs.2015.0053 variation in Darwin’s finches. Proc. Natl. Acad. Sci. USA 107: Draghi, J., and G. P. Wagner, 2009 The evolutionary dynamics of 3356–3360. https://doi.org/10.1073/pnas.0911575107 evolvability in a gene network model. J. Evol. Biol. 22: 599–611. Casasa, S., and A. P. Moczek, 2018 The role of ancestral pheno- https://doi.org/10.1111/j.1420-9101.2008.01663.x typic plasticity in evolutionary diversification: population den- Draghi, J. A., and M. C. Whitlock, 2012 Phenotypic plasticity fa- sity effects in horned beetles. Anim. Behav. 137: 53–61. cilitates mutational variance, genetic variance, and evolvability Cavalli-Sforza, L. L., and M. W. Feldman, 1981 Cultural Trans- along the major axis of environmental variation. Evolution 66: mission and Evolution. Princeton University Press, Princeton, NJ. 2891–2902. https://doi.org/10.1111/j.1558-5646.2012.01649.x Charlesworth, B., 1976 Recombination modification in a fluctuat- Erwin, D. H., 2017 Developmental push or environmental pull? ing environment. Genetics 83: 181–195. The causes of macroevolutionary dynamics. Hist. Philos. Life Sci. Charlesworth, B., and R. Lande, 1982 Morphological stasis and 39: 36. https://doi.org/10.1007/s40656-017-0163-0 developmental constraint: no problem for Neo-Darwinism. Na- Espinosa-Soto, C., O. C. Martin, and A. Wagner, 2011 Phenotypic ture 296: 610. https://doi.org/10.1038/296610a0 plasticity can facilitate adaptive evolution in gene regulatory circuits. Charlesworth, B., R. Lande, and M. Slatkin, 1982 A Neo-Darwinian BMC Evol. Biol. 11: 5. https://doi.org/10.1186/1471-2148-11-5 commentary on macroevolution. Evolution 36: 474–498. https:// Evans, A. R., E. S. Daly, K. K. Catlett, K. S. Paul, S. J. King et al., doi.org/10.1111/j.1558-5646.1982.tb05068.x 2016 A simple rule governs the evolution and development of Charlesworth, D., N. H. Barton, and B. Charlesworth, 2017 The hominin tooth size. Nature 530: 477–480. https://doi.org/ sources of adaptive variation. Proc. Biol. Sci. 284: 20162864. 10.1038/nature16972 https://doi.org/10.1098/rspb.2016.2864 Falconer, D. S., and T. F. C. Mackay, 1996 Introduction to Quan- Chebib, J., and F. Guillaume, 2017 What affects the predictability titative Genetics. Prentice Hall, Essex, UK. of evolutionary constraints using a G-matrix? The relative ef- Feldman, M. W., and U. Liberman, 1986 An evolutionary reduc- fects of modular pleiotropy and mutational correlation. Evolu- tion principle for genetic modifiers. Proc. Natl. Acad. Sci. USA tion 71: 2298–2312. https://doi.org/10.1111/evo.13320 83: 4824–4827. https://doi.org/10.1073/pnas.83.13.4824 Cheverud, J., 2017 Genetics: role of mutation in fly-wing evolution. Félix, M. A., 2016 Phenotypic evolution with and beyond genome Nature 548: 401–403. https://doi.org/10.1038/nature23536 evolution. Curr. Top. Dev. Biol. 119: 291–347. https://doi.org/ Cheverud, J. M., 1984 Quantitative genetics and developmental 10.1016/bs.ctdb.2016.04.002 constraints on evolution by selection. J. Theor. Biol. 110: 155– Félix, M. A., and M. Barkoulas, 2012 Robustness and flexibility in 171. https://doi.org/10.1016/S0022-5193(84)80050-8 nematode vulva development. Trends Genet. 28: 185–195. Cheverud, J. M., 1996 Developmental integration and the evolu- https://doi.org/10.1016/j.tig.2012.01.002 tion of pleiotropy. Am. Zool. 36: 44–50. https://doi.org/10.1093/ Fierst, J. L., 2011 A history of phenotypic plasticity accelerates icb/36.1.44 adaptation to a new environment. J. Evol. Biol. 24: 1992– Ciliberti, S., O. C. Martin, and A. Wagner, 2007 Innovation and 2001. https://doi.org/10.1111/j.1420-9101.2011.02333.x robustness in complex regulatory gene networks. Proc. Natl. Fontana, W., and P. Schuster, 1998 Continuity in evolution: on Acad. Sci. USA 104: 13591–13596. https://doi.org/10.1073/ the nature of transitions. Science 280: 1451–1455. https://doi. pnas.0705396104 org/10.1126/science.280.5368.1451 Clark, E., 2017 Dynamic patterning by the Drosophila pair-rule net- Frank, S. A., 2011 Natural selection. II. Developmental variability work reconciles long-germ and short-germ segmentation. PLoS and evolutionary rate*. J. Evol. Biol. 24: 2310–2320. https:// Biol. 15: e2002439. https://doi.org/10.1371/journal.pbio.2002439 doi.org/10.1111/j.1420-9101.2011.02373.x

962 T. Uller et al. Fritz, J. A., J. Brancale, M. Tokita, K. J. Burns, M. B. Hawkins et al., Hughes, M., S. Gerber, and M. A. Wills, 2013 Clades reach highest 2014 Shared developmental programme strongly constrains morphological disparity early in their evolution. Proc. Natl. beak shape diversity in songbirds. Nat. Commun. 5: 3700. Acad. Sci. USA 110: 13875–13879. https://doi.org/10.1073/ https://doi.org/10.1038/ncomms4700 pnas.1302642110 Futuyma, D. J., 2015 Can modern evolutionary theory explain Jablonski, D., 2017 Approaches to macroevolution: 1. General macroevolution? in Macroevolution: Explanation, Interpretation concepts and origin of variation. Evol. Biol. 44: 427–450. https:// and Evidence, edited by E. Serrelli, and N. Gontier. Springer In- doi.org/10.1007/s11692-017-9420-0 ternational Publishing, Basel, Switzerland. https://doi.org/ Jaeger, J., and A. Crombach, 2012 Life’s attractors understanding 10.1007/978-3-319-15045-1_2 developmental systems through reverse engineering and Futuyma, D. J., 2017 Evolutionary biology today and the call for in silico evolution, pp. 93–119 in Evolutionary Systems Biology, an extended synthesis. Interface Focus 7: 20160145. https:// edited by O. S. Soyer. Springer, New York. https://doi.org/ doi.org/10.1098/rsfs.2016.0145 10.1007/978-1-4614-3567-9_5 Galis, F., 1999 Why do almost all mammals have seven cervical Jaeger, J., and N. Monk, 2014 Bioattractors: dynamical systems vertebrae? Developmental constraints, Hox genes, and cancer. theory and the evolution of regulatory processes. J. Physiol. J. Exp. Zool. 285: 19–26. https://doi.org/10.1002/(SICI)1097- 592: 2267–2281. https://doi.org/10.1113/jphysiol.2014.272385 010X(19990415)285:1,19::AID-JEZ3.3.0.CO;2-Z Jiménez,A.,J.Cotterell,A.Munteanu,andJ.Sharpe,2015 Dynamics Galis, F., T. J. M. Van Dooren, J. D. Feuth, J. A. J. Metz, A. Witkam of gene circuits shapes evolvability. Proc. Natl. Acad. Sci. USA 112: et al., 2006 Extreme selection in humans against homeotic 2103–2108 (erratum: Proc. Natl. Acad. Sci. USA 112: E5110). transformations of cervical vertebrae. Evolution 60: 2643– https://doi.org/10.1073/pnas.1411065112 2654. https://doi.org/10.1111/j.0014-3820.2006.tb01896.x Jones, A. G., S. J. Arnold, and R. Burger, 2007 The mutation Geoghegan, J. L., and H. G. Spencer, 2012 Population-epigenetic matrix and the evolution of evolvability. Evolution 61: 727– models of selection. Theor. Popul. Biol. 81: 232–242. https:// 745. https://doi.org/10.1111/j.1558-5646.2007.00071.x doi.org/10.1016/j.tpb.2011.08.001 Jones, A. G., R. Burger, and S. J. Arnold, 2014 Epistasis and natural Gerhart, J., and M. Kirschner, 2007 The theory of facilitated var- selection shape the mutational architecture of complex traits. Nat. iation. Proc. Natl. Acad. Sci. USA 104: 8582–8589. https://doi. Commun. 5: 3709. https://doi.org/10.1038/ncomms4709 org/10.1073/pnas.0701035104 Kashtan, N., and U. Alon, 2005 Spontaneous evolution of modu- Gomez-Mestre, I., and D. R. Buchholz, 2006 Developmental plas- larity and network motifs. Proc. Natl. Acad. Sci. USA 102: ticity mirrors differences among taxa in spadefoot toads linking 13773–13778. https://doi.org/10.1073/pnas.0503610102 plasticity and diversity. Proc. Natl. Acad. Sci. USA 103: 19021– Kashtan, N., E. Noor, and U. Alon, 2007 Varying environments 19026. https://doi.org/10.1073/pnas.0603562103 can speed up evolution. Proc. Natl. Acad. Sci. USA 104: 13711– Goswami, A., W. J. Binder, J. Meachen, and F. R. O’Keefe, 13716. https://doi.org/10.1073/pnas.0611630104 2015 The fossil record of phenotypic integration and modular- Kauffman, S., 1969 Homeostasis and differentiation in random ity: a deep-time perspective on developmental and evolutionary genetic control networks. Nature 224: 177–178. https://doi. dynamics. Proc. Natl. Acad. Sci. USA 112: 4891–4896. https:// org/10.1038/224177a0 doi.org/10.1073/pnas.1403667112 Kauffman, S. A., 1983 Developmental constraints: internal factors Gromko, M. H., 1995 Unpredictability of correlated response to in evolution in Development and Evolution, edited by B. C. Good- selection: pleiotropy and sampling interact. Evolution 49: 685– win, N. Holder, and C. C. Wylie. Cambridge University Press, 693. https://doi.org/10.1111/j.1558-5646.1995.tb02305.x Cambridge, UK. Hall, B. K., 2015 Bones and Cartilage: Developmental and Evolu- Kavanagh, K. D., A. R. Evans, and J. Jernvall, 2007 Predicting tionary Skeletal Biology. Elsevier, London. evolutionary patterns of mammalian teeth from development. Hansen, T. F., and D. Houle, 2008 Measuring and comparing evolv- Nature 449: 427–432. https://doi.org/10.1038/nature06153 ability and constraint in multivariate characters. J. Evol. Biol. 21: Kavanagh, K. D., O. Shoval, B. B. Winslow, U. Alon, B. P. Leary 1201–1219. https://doi.org/10.1111/j.1420-9101.2008.01573.x et al., 2013 Developmental bias in the evolution of phalanges. Hansen, T. F., J. M. Alvarez-Castro, A. J. R. Carter, J. Hermisson, Proc. Natl. Acad. Sci. USA 110: 18190–18195. https://doi.org/ and G. P. Wagner, 2006 Evolution of genetic architecture un- 10.1073/pnas.1315213110 der directional selection. Evolution 60: 1523–1536. https://doi. Kiontke, K., A. Barriere, I. Kolotuev, B. Podbilewicz, R. Sommer org/10.1111/j.0014-3820.2006.tb00498.x et al., 2007 Trends, stasis, and drift in the evolution of nema- Harjunmaa, E., K. Seidel, T. Hakkinen, E. Renvoise, I. J. Corfe et al., tode vulva development. Curr. Biol. 17: 1925–1937. https://doi. 2014 Replaying evolutionary transitions from the dental org/10.1016/j.cub.2007.10.061 fossil record. Nature 512: 44–48. https://doi.org/10.1038/ Kirschner, M. W., and J. C. Gerhart, 2005 The Plausibility of Life: nature13613 Resolving Darwin’s Dilemma. Yale University Press, New Haven, CT. Held, L., 2017 Deep ? Uncanny Similarities of Humans Klingenberg, C. P., 2014 Studying morphological integration and and Flies Uncovered by Evo-Devo. Cambridge University Press, modularity at multiple levels: concepts and analysis. Philos. Cambridge, UK. https://doi.org/10.1017/9781316550175 Trans. R. Soc. Lond. B Biol. Sci. 369: 20130249. https://doi. Hinton, G. E., and S. J. Nowlan, 1987 How learning can guide org/10.1098/rstb.2013.0249 evolution. Complex Syst. 1: 495–502. Klingenberg, C. P., and S. D. Zaklan, 2000 Morphological integra- Houle, D., 1991 Genetic covariance of fitness correlates: what tion between developmental compartments in the Drosophila genetic correlations are made of and why it matters. Evolution 45: wing. Evolution 54: 1273–1285. https://doi.org/10.1111/j.0014- 630–648. https://doi.org/10.1111/j.1558-5646.1991.tb04334.x 3820.2000.tb00560.x Houle, D., and J. Fierst, 2013 Properties of spontaneous muta- Kounios, L., J. Clune, K. Kouvaris, G. P. Wagner, M. Pavlicev et al., tional variance and covariance for wing size and shape in Dro- 2017 Resolving the paradox of evolvability with learning the- sophila melanogaster. Evolution 67: 1116–1130. https://doi. ory: how evolution learns to improve evolvability on rugged org/10.1111/j.1558-5646.2012.01838.x fitness landscapes. arXiv: 1612.05955. Houle, D., G. H. Bolstad, K. van der Linde, and T. F. Hansen, Kouvaris, K., J. Clune, L. Kounios, M. Brede, and R. A. Watson, 2017 Mutation predicts 40 million years of fly wing evolution. 2017 How evolution learns to generalise: using the principles Nature 548: 447–450. https://doi.org/10.1038/nature23473 of learning theory to understand the evolution of developmental

Developmental Bias and Evolution 963 organisation. PLoS Comput. Biol. 13: e1005358. https://doi. Rev. Ecol. Evol. Syst. 47: 463–486. https://doi.org/10.1146/ org/10.1371/journal.pcbi.1005358 annurev-ecolsys-121415-032409 Kovaka, K., 2017 Underdetermination and evidence in the devel- Mercader, N., E. Leonardo, N. Azpiazu, A. Serrano, G. Morata et al., opmental plasticity debate. Br. J. Philos. Sci., axx038. https:// 1999 Conserved regulation of proximodistal limb axis devel- doi.org/10.1093/bjps/axx038 opment by Meis1/Hth. Nature 402: 425–429. https://doi.org/ Kozmik, Z., 2005 Pax genes in eye development and evolution. 10.1038/46580 Curr. Opin. Genet. Dev. 15: 430–438. https://doi.org/10.1016/ Mezey, J. G., and D. Houle, 2005 The dimensionality of genetic j.gde.2005.05.001 variation for wing shape in Drosophila melanogaster. Evolution Kozmik, Z., J. Ruzickova, K. Jonasova, Y. Matsumoto, P. Vopalensky 59: 1027–1038. https://doi.org/10.1111/j.0014-3820.2005. et al., 2008 Assembly of the cnidarian camera-type eye from tb01041.x vertebrate-like components. Proc. Natl. Acad. Sci. USA 105: Mills, R., and R. A. Watson, 2006 On crossing fitness valleys with 8989–8993. https://doi.org/10.1073/pnas.0800388105 the Baldwin effect, pp. 493–499 in Proceedings of the Tenth In- Kulkarni, S. S., R. J. Denver, I. Gomez-Mestre, and D. R. Buchholz, ternational Conference on the Simulation and Synthesis of Living 2017 Genetic accommodation via modified endocrine signal- Systems. MIT Press, Cambridge, MA. ling explains phenotypic divergence among spadefoot toad Moczek, A. P., S. Sultan, S. Foster, C. Ledon-Rettig, I. Dworkin et al., species. Nat. Commun. 8: 993. https://doi.org/10.1038/ 2011 The role of developmental plasticity in evolutionary in- s41467-017-00996-5 novation. Proc. Biol. Sci. 278: 2705–2713. https://doi.org/ Lande, R., 1980 The genetic covariance between characters main- 10.1098/rspb.2011.0971 tained by pleiotropic mutations. Genetics 94: 203–215. Morrissey, M. B., 2015 Evolutionary quantitative genetics of non- Lande, R., and S. J. Arnold, 1983 The measurement of selection linear developmental systems. Evolution 69: 2050–2066. https:// on correlated characters. Evolution 37: 1210–1226. https://doi. doi.org/10.1111/evo.12728 org/10.1111/j.1558-5646.1983.tb00236.x Muller, H. J., 1922 Variation due to change in the individual gene. Levis, N. A., and D. W. Pfennig, 2016 Evaluating ‘Plasticity-First’ Am. Nat. 56: 32–50. https://doi.org/10.1086/279846 evolution in nature: key criteria and empirical approaches. Muschick, M., M. Barluenga, W. Salzburger, and A. Meyer, Trends Ecol. Evol. 31: 563–574. https://doi.org/10.1016/j. 2011 Adaptive phenotypic plasticity in the Midas cichlid fish tree.2016.03.012 pharyngeal jaw and its relevance in adaptive radiation. BMC Lipson, H., J. B. Pollack, and N. P. Suh, 2002 On the origin of Evol. Biol. 11: 116. https://doi.org/10.1186/1471-2148-11-116 modular variation. Evolution 56: 1549–1556. https://doi.org/ Nakamura, T., A. R. Gehrke, J. Lemberg, J. S. Zymaszek, and N. H. 10.1111/j.0014-3820.2002.tb01466.x Shubin, 2016 Digits and fin rays share common developmen- Losos, J. B., 2010 Adaptive radiation, ecological opportunity, and tal histories. Nature 537: 225–228. https://doi.org/10.1038/ evolutionary determinism. Am. Nat. 175: 623–639. https://doi. nature19322 org/10.1086/652433 Nei, M., 2013 Mutation-Driven Evolution. Oxford University Press, Losos, J. B., 2011 Convergence, adaptation, and constraint. Evo- Oxford. lution 65: 1827–1840. https://doi.org/10.1111/j.1558-5646.2011. Newman, S. A., and G. B. Müller, 2005 Origination and innova- 01289.x tion in the vertebrate limb skeleton: an epigenetic perspective. Lynch, M. W. B., 1998 Genetics and Analysis of Quantitative Traits, J. Exp. Zoolog. B Mol. Dev. Evol. 304: 593–609. https://doi.org/ Sinauer, Sunderland, MA. 10.1002/jez.b.21066 Lynch, M., and B. Walsh, 1998 Genetics and Analysis of Quantita- Olson, E. N., 2006 Gene regulatory networks in the evolution and tive Traits, Sinauer, Sunderland, MA. development of the heart. Science 313: 1922–1927. https://doi. Mallarino, R., P. R. Grant, B. R. Grant, A. Herrel, W. P. Kuo et al., org/10.1126/science.1132292 2011 Two developmental modules establish 3D beak-shape Oster, G., and P. Alberch, 1982 Evolution and bifurcation of de- variation in Darwin’s finches. Proc. Natl. Acad. Sci. USA 108: velopmental programs. Evolution 36: 444–459. https://doi.org/ 4057–4062. https://doi.org/10.1073/pnas.1011480108 10.1111/j.1558-5646.1982.tb05066.x Matamoro-Vidal, A., I. Salazar-Ciudad, and D. Houle, 2015 Making Oyston, J. W., M. Hughes, S. Gerber, and M. A. Wills, 2016 Why quantitative morphological variation from basic developmental should we investigate the morphological disparity of plant processes: where are we? The case of the Drosophila wing. Dev. clades? Ann. Bot. (Lond.) 117: 859–879. https://doi.org/10.1093/ Dyn. 244: 1058–1073. https://doi.org/10.1002/dvdy.24255 aob/mcv135 Maynard-Smith, J., R. Burian, S. Kauffman, P. Alberch, J. Campbell Paenke, I., B. Sendhoff, and T. J. Kawecki, 2007 Influence of et al., 1985 Developmental constraints and evolution. Q. Rev. plasticity and learning on evolution under directional selection. Biol. 60: 265–287. https://doi.org/10.1086/414425 Am. Nat. 170: E47–E58. https://doi.org/10.1086/518952 McDonald, M. J., S. M. Gehrig, P. L. Meintjes, X. X. Zhang, and P. B. Panganiban, G., S. M. Irvine, C. Lowe, H. Roehl, L. S. Corley et al., Rainey, 2009 Adaptive divergence in experimental populations 1997 The origin and evolution of animal appendages. Proc. of Pseudomonas fluorescens. IV. Genetic constraints guide evo- Natl. Acad. Sci. USA 94: 5162–5166. https://doi.org/10.1073/ lutionary trajectories in a parallel adaptive radiation. Genetics pnas.94.10.5162 183: 1041–1053. https://doi.org/10.1534/genetics.109.107110 Parter, M., N. Kashtan, and U. Alon, 2008 Facilitated variation: McGhee, G., 2007 The Geometry of Evolution. Adaptive Landscapes how evolution learns from past environments to generalize to and Theoretical Morphospaces. Cambridge University Press, New new environments. PLoS Comput. Biol. 4: e1000206. https:// York. doi.org/10.1371/journal.pcbi.1000206 McNamara, J. M., S. R. X. Dall, P. Hammerstein, and O. Leimar, Pavlicev, M., J. M. Cheverud, and G. P. Wagner, 2011 Evolution of 2016 Detection vs. selection: integration of genetic, epigenetic adaptive phenotypic variation patterns by direct selection for and environmental cues in fluctuating environments. Ecol. Lett. evolvability. Proc. Biol. Sci. 278: 1903–1912. https://doi.org/ 19: 1267–1276. https://doi.org/10.1111/ele.12663 10.1098/rspb.2010.2113 McShea, D. W., 1994 Mechanisms of large-scale evolutionary Payne, J. L., J. H. Moore, and A. Wagner, 2014 Robustness, evolv- trends. Evolution 48: 1747–1763. https://doi.org/10.1111/ ability, and the logic of genetic regulation. Artif. Life 20: 111– j.1558-5646.1994.tb02211.x 126. https://doi.org/10.1162/ARTL_a_00099 Melo, D., A. Porto, J. M. Cheverud, and G. Marroig, Peter, I. S., and E. H. Davidson, 2015 Genomic Control Processes. 2016 Modularity: genes, development, and evolution. Annu. Development and Evolution. Academic Press, Oxford.

964 T. Uller et al. Pfennig, D. W., M. A. Wund, E. C. Snell-Rood, T. Cruickshank, C. D. in a model organism from weaning to adulthood. J. Exp. Biol. Schlichting et al., 2010 Phenotypic plasticity’s impacts on di- 217: 4099–4107. versification and . Trends Ecol. Evol. 25: 459–467. Scoville, A. G., and M. E. Pfrender, 2010 Phenotypic plasticity https://doi.org/10.1016/j.tree.2010.05.006 facilitates recurrent rapid adaptation to introduced predators. Pie, M. R., and J. S. Weitz, 2005 A null model of morphospace occu- Proc. Natl. Acad. Sci. USA 107: 4260–4263. https://doi.org/ pation. Am. Nat. 166: E1–E13. https://doi.org/10.1086/430727 10.1073/pnas.0912748107 Prud’homme,B.,N.Gompel,A.Rokas,V.A.Kassner,T.M.Williamset al., Shubin, N., C. Tabin, and S. Carroll, 2009 Deep homology and the 2006 Repeated morphological evolution through cis-regulatory origins of evolutionary novelty. Nature 457: 818–823. https:// changes in a pleiotropic gene. Nature 440: 1050–1053. https:// doi.org/10.1038/nature07891 doi.org/10.1038/nature04597 Simon, M. N., F. A. Machado, and G. Marroig, 2016 High evolu- Prud’homme, B., N. Gompel, and S. B. Carroll, 2007 Emerging tionary constraints limited adaptive responses to past climate principles of regulatory evolution. Proc. Natl. Acad. Sci. USA changes in toad skulls. Proc. Biol. Sci. B 283: 20161783. DOI: 104: 8605–8612. https://doi.org/10.1073/pnas.0700488104 10.1098/rspb.2016.1783. Psujek, S., and R. D. Beer, 2008 Developmental bias in evolution: Snell-Rood, E. C., 2012 Selective processes in development: im- evolutionary accessibility of phenotypes in a model evo-devo plications for the costs and benefits of phenotypic plasticity. In- system. Evol. Dev. 10: 375–390. https://doi.org/10.1111/ tegr. Comp. Biol. 52: 31–42. https://doi.org/10.1093/icb/ j.1525-142X.2008.00245.x ics067 Rabosky, D. L., 2009 Ecological limits on clade diversification in Standen, E. M., T. Y. Du, and H. C. E. Larsson, 2014 Developmental higher taxa. Am. Nat. 173: 662–674. https://doi.org/10.1086/ plasticity and the origin of tetrapods. Nature 513: 54–58. https:// 597378 doi.org/10.1038/nature13708 Raup, D. M., 1966 Geometric analysis of shell coiling: general Stoltzfus, A., and D. M. McCandlish, 2017 Mutational biases in- problems. J. Paleontol. 40: 1178–1190. fluence parallel adaptation. Mol. Biol. Evol. 34: 2163–2172. Rice, S. A., 2004 Evolutionary Theory. Mathematical and Concep- https://doi.org/10.1093/molbev/msx180 tual Foundations. Sinauer, Sunderland, MA. Suzuki, Y., and H. F. Nijhout, 2006 Evolution of a polyphenism by Rice, S. H., 2008 Theoretical approaches to the evolution of de- genetic accommodation. Science 311: 650–652. https://doi. velopment and genetic architecture. Ann. N. Y. Acad. Sci. 1133: org/10.1126/science.1118888 67–86. https://doi.org/10.1196/annals.1438.002 Todd, P. M. G., 1991 Exploring adaptive agency II: simulating the Salazar-Ciudad, I., 2006 Developmental constraints vs. varia- evolution of associative learning, pp. 306–315 in From Animals tional properties: how can help to understand to Animals: Proceedings of the First International Conference on evolution and development. J. Exp. Zoolog. B Mol. Dev. Evol. Simulation of Adaptive Behavior, edited by J. M. S. Wilson. MIT 306B: 107–125. https://doi.org/10.1002/jez.b.21078 Press, Cambridge, MA. Salazar-Ciudad, I., 2008 Evolution in biological and nonbiological van Gestel, J., and F. J. Weissing, 2016 Regulatory mechanisms systems under different mechanisms of generation and inheri- link phenotypic plasticity to evolvability. Sci. Rep. 6: 24524. tance. Theory Biosci. 127: 343–358. https://doi.org/10.1007/ https://doi.org/10.1038/srep24524 s12064-008-0052-x von Dassow, G., E. Meir, E. M. Munro, and G. M. Odell, 2000 The Salazar-Ciudad, I., and J. Jernvall, 2005 Graduality and innova- segment polarity network is a robust development module. Na- tion in the evolution of complex phenotypes: insights from de- ture 406: 188–192. https://doi.org/10.1038/35018085 velopment. J. Exp. Zoolog. B Mol. Dev. Evol. 304: 619–631. Waddington, C. H., 1957 Strategy of the Genes. George Allen & https://doi.org/10.1002/jez.b.21058 Unwin Ltd, London. Salazar-Ciudad, I., and J. Jernvall, 2010 A computational model of Wagner, A., 2011 The Origins of Evolutionary Innovations. Oxford teeth and the developmental origins of morphological variation. University Press, Oxford. https://doi.org/10.1093/acprof:oso/ Nature 464: 583–586. https://doi.org/10.1038/nature08838 9780199692590.001.0001 Salazar-Ciudad, I., J. Jernvall, and S. A. Newman, 2003 Mechanisms Wagner, G., 2014 Homology, Genes, and Evolutionary Innovation. of pattern formation in development and evolution. Development Princeton University Press, Princeton, NJ. https://doi.org/ 130: 2027–2037. https://doi.org/10.1242/dev.00425 10.1515/9781400851461 Sanger, T. J., E. A. Norgard, L. S. Pletscher, M. Bevilacqua, V. R. Wagner, G. P., and L. Altenberg, 1996 Perspective: complex ad- Brooks et al., 2011 Developmental and genetic origins of mu- aptations and the evolution of evolvability. Evolution 50: 967– rine long bone length variation. J. Exp. Zoolog. B Mol. Dev. Evol. 976. https://doi.org/10.1111/j.1558-5646.1996.tb02339.x 316B: 146–161. https://doi.org/10.1002/jez.b.21388 Wake, D. B., 1991 Homoplasy - the result of natural-selection, or Scharloo, W., 1970 Stabilizing and disruptive selection on a evidence of design limitations. Am. Nat. 138: 543–567. https:// mutant character in Drosophila. 3. Polymorphism caused doi.org/10.1086/285234 by a developmental switch mechanism. Genetics 65: 693– Watson, R. A., and E. Szathmary, 2016 How can evolution learn? 705. Trends Ecol. Evol. 31: 147–157. https://doi.org/10.1016/j. Schlosser, G., 2002 Modularity and the units of evolution. Theory tree.2015.11.009 Biosci. 121: 1–80. https://doi.org/10.1078/1431-7613-00049 Watson, R. A., G. P. Wagner, M. Pavlicev, D. M. Weinreich, and R. Schluter, D., 1996 Adaptive radiation along genetic lines of least Mills, 2014 The evolution of phenotypic correlations and “de- resistance. Evolution 50: 1766–1774. https://doi.org/10.1111/ velopmental memory”. Evolution 68: 1124–1138. https://doi. j.1558-5646.1996.tb03563.x org/10.1111/evo.12337 Schluter, D., 2000 The Ecology of Adaptive Radiation. Oxford Uni- Watson,R.A.,R.Mills,C.L.Buckley,K.Kouvaris,A.Jackson versity Press, Oxford. et al., 2016 Evolutionary connectionism: algorithmic prin- Schmalhausen, I. I., 1949 Factors of Evolution. The Theory of Sta- ciples underlying the evolution of biological organisation bilizing Selection. Blakiston, Philadelphia, PA. in evo-devo, evo-eco and evolutionary transitions. Evol. Schneider, R. F., and A. Meyer, 2017 How plasticity, genetic as- Biol. 43: 553–581. https://doi.org/10.1007/s11692-015- similation and cryptic genetic variation may contribute to adap- 9358-z tive radiations. Mol. Ecol. 26: 330–350. Wessinger, C. A., and L. C. Hileman, 2016 Accessibility, constraint, Scott, J. E., K. R. McAbee, M. M. Eastman, and M. J. Ravosa, and repetition in adaptive floral evolution. Dev. Biol. 419: 175– 2014 Teaching an old jaw new tricks: diet-induced plasticity 183. https://doi.org/10.1016/j.ydbio.2016.05.003

Developmental Bias and Evolution 965 West-Eberhard, M. J., 2003 Developmental Plasticity and Evolu- Wund, M. A., J. A. Baker, B. Clancy, J. L. Golub, and S. A. Fosterk, tion. Oxford University Press, New York. 2008 A test of the “Flexible stem” model of evolution: ances- Wilkins, A. S., 2005 Recasting developmental evolution in terms tral plasticity, genetic accommodation, and morphological diver- of genetic pathway and network evolution ... and the implica- gence in the threespine stickleback radiation. Am. Nat. 172: tions for comparative biology. Brain Res. Bull. 66: 495–509. 449–462. https://doi.org/10.1086/590966 https://doi.org/10.1016/j.brainresbull.2005.04.001 Xavier-Neto, J., R. A. Castro, A. C. Sampaio, A. P. Azambuja, H. A. Wilkins, A. S., 2007 Between “design” and “bricolage”: genetic Castillo et al., 2007 Parallel avenues in the evolution of hearts networks, levels of selection, and adaptive evolution. Proc. Natl. and pumping organs. Cell. Mol. Life Sci. 64: 719–734. https:// Acad. Sci. USA 104: 8590–8596. https://doi.org/10.1073/ doi.org/10.1007/s00018-007-6524-1 pnas.0701044104 Yampolsky, L. Y., and A. Stoltzfus, 2001 Bias in the introduction Wittkopp, P. J., and G. Kalay, 2012 Cis-regulatory elements: molecu- of variation as an orienting factor in evolution. Evol. Dev. 3: 73– lar mechanisms and evolutionary processes underlying divergence. 83. https://doi.org/10.1046/j.1525-142x.2001.003002073.x Nat. Rev. Genet. 13: 59–69. https://doi.org/10.1038/nrg3095 Young, R. L., and A. V. Badyaev, 2010 Developmental plasticity Wotton, K. R., E. Jimenez-Guri, A. Crombach, H. Janssens, A. Alcaine- links local adaptation and evolutionary diversification in forag- Colet et al., 2015 Quantitative system drift compensates for al- ing morphology. J. Exp. Zoolog. B Mol. Dev. Evol. 314: 434–444. tered maternal inputs to the gap gene network of the scuttle fly https://doi.org/10.1002/jez.b.21349 Megaselia abdita. eLife 4: e04785. https://doi.org/10.7554/ Young, R. L., M. J. Sweeney, and A. V. Badyaev, 2010 Morphological eLife.04785 diversity and ecological similarity: versatility of muscular and skel- Wray,G.A.,2007 Theevolutionarysignificance of cis-regulatory muta- etal morphologies enables ecological convergence in shrews. tions. Nat. Rev. Genet. 8: 206–216. https://doi.org/10.1038/nrg2063 Funct. Ecol. 24: 556–565. https://doi.org/10.1111/j.1365-2435.2009. Wright, D. F., 2017 Phenotypic innovation and adaptive con- 01664.x straints in the evolutionary radiation of palaeozoic crinoids. Sci. Rep. 7: 13745. https://doi.org/10.1038/s41598-017-13979-9 Communicating editor: A. S. Wilkins

966 T. Uller et al.