C HAPTER 10

PHYLOGENETIC APPROACHES FOR RESEARCH IN COMPARATIVE COGNITION

Evan L. MacLean and Charles L. Nunn

Comparative studies have the potential to address be treated as independent observationsASSOCIATION in statisti- a wide range of questions about how, when, and cal analyses because of patterns of inheritance on a why different traits have evolved. The compara- bifurcating phylogeny (Felsenstein, 1985; Martins & tive approach refers to studies in which variation in Garland, 1991). To deal with this nonindependence, traits of different (or different populations) comparative biologists incorporate information is used to test specifc hypotheses or to generate new about the evolutionary relationships between species. hypotheses about evolutionary phenomena. In many Phylogenetic comparative methods are a set of sta- cases, comparative methods are used to investigate tistical approaches designed for exactly this purpose how two or more traits covary. Comparative meth- (Garamszegi,PSYCHOLOGICAL 2014; Garland, Bennett, & Rezende, ods are also used to reconstruct evolutionary history 2005; Nunn, 2011; Rezende & Diniz-Filho, 2012), or to assess how traits infuence patterns of diver- a toolkit that is becoming increasingly useful in the sifcation (speciation and extinction; Nunn, 2011). feld of comparative psychology. In addition to being In the context of comparative psychology, these applied in a broader range of evolutionary contexts, methods have been used, for example, to investigate comparative methods themselves are evolving, and associations between aspects of neuroanatomyAMERICAN and many classical techniques are rapidly being replaced executive function (MacLean et al., 2014;© Shultz & by newer and more fexible approaches. Dunbar, 2010; see also Chapter 24, this volume), In this chapter, we introduce the reader to a to explore how life history traits infuence temporal range of phylogenetic comparative methods that decision making (Stevens, 2014; see also Volume 2, can be used to address fundamental questions in Chapter 24, this handbook),PROOFS and to make inferences comparative psychology, including some recent about how life in complex societies has shaped pri- methodological advances that will create powerful mate cognitive evolution (Amici, Aureli, & Call, opportunities for future research. We illustrate these 2008; Burkart et al., 2014; MacLean et al., 2013; see methods by using a combination of simulated data also Chapters 12 and 13, this volume). and analyses of published datasets. Given the rapid Understanding evolutionary relationships—or growth of research on comparative cognition, we phylogeny—among the species of interest is critical highlight the utility of comparative methods for the for effective inference in comparative research. In a study of cognitive evolution. However, the concepts UNCORRECTEDvery real sense, phylogeny is the scaffolding on which and statistical approaches described in this chapter one investigates evolutionary change in traits and the can be similarly applied in other areas of compara- factors that lead to these changes. This is important tive psychology. conceptually, and it is also critically relevant in statis- At the outset, we want to emphasize that tical analyses. For example, in the context of studying some traits are measured on a quasi-continuous how traits covary, data on different species cannot scale (e.g., percentage of correct responses on an

http://dx.doi.org/10.1037/XXXXX-XXX APA Handbook of Comparative Psychology: Vol. 1. Basic Concepts, Methods, Neural Substrate, and Behavior, J. Call (Editor-in-Chief) 1 Copyright © 2017 by the American Psychological Association. All rights reserved.

BK-APA-HCM_V1-160213-Chp10.indd 1 04/07/16 6:13 PM MacLean and Nunn

experimental task), and other traits are measured terms and concepts that are foundational to all phy- on a discrete scale (e.g., presence or absence of mir- logenetic comparative methods. ror self-recognition). These two types of data often require different statistical approaches. For brevity, PHYLOGENETIC TREES we present a single variant of each method below, yet we also direct the reader to relevant literature on As just noted, phylogenies represent the evolutionary other approaches throughout and to several recent relationships between taxa and are frequently reviews (Garamszegi, 2014; Garland et al., 2005; visualized as trees with a branching pattern (see Nunn, 2011; Rezende & Diniz-Filho, 2012). We Figure 10.1A). Phylogenetic trees consist of nodes and begin with a brief introduction to additional key branches. Nodes indicate speciation events where an

(A)

ASSOCIATION

PSYCHOLOGICAL

(B) ABCDEFGH

A41000000 B 1 4 AMERICAN000000 C 00© 4 32111

D 0034 2 111

E 0022 4 111

F PROOFS00111 4 32

G 001113 4 2

H 00111224

FIGURE 10.1. A phylogenetic tree and variance–covariance matrix representing the evolutionary relationships between species. A: The root of the tree is at the bottom with the tips (terminal branches) extend- UNCORRECTEDing upward. Uppercase letters refer to extent taxa at the tips of the tree. Nodes are indicated by lowercase letters enclosed in circles, and branch lengths are shown to the left of each branch in the tree. The scale bar indicates 1 million years (MY). B: The variance–covariance matrix repre- sents the total age of the phylogeny along the diagonal, with the extent of shared evolutionary history between pairs of species indicated on the off-diagonals.

2

BK-APA-HCM_V1-160213-Chp10.indd 2 04/07/16 6:13 PM Phylogenetic Approaches for Research in Comparative Cognition

ancestral lineage gave rise to two (or more) descendent PHYLOGENETIC SIGNAL species. Nodes are connected by branches, which are typically drawn to be proportional to evolutionary As a result of shared evolutionary history, closely time. Figure 10.1A shows a phylogenetic tree for eight related species tend to resemble one another more so species with circles enclosing the internal nodes. Node than less closely related taxa; this tendency is termed r is located at the root of the phylogeny and repre- phylogenetic signal (Blomberg & Garland, 2002; sents the oldest bifurcation in the tree, which in this Blomberg, Garland, & Ives, 2003). Interestingly, the example occurred 4 million years ago. The internal extent to which traits are associated with phylogeny branches of the tree (i.e., branches connecting nodes, varies widely from one trait to another, with morpho- rather than ending at a tip) represent the time that logical traits tending to exhibit higher levels of phylo- species have shared evolutionary history, whereas the genetic signal than behavioral, cognitive, or ecological terminal branches (leading to a tip) represent the time variables (Blomberg et al., 2003; Kamilar & Cooper, that each extant species has evolved independently of 2013; MacLean et al., 2012). Quantitative estimates of phylogenetic signal in continuous traits can be other taxa in the tree (of course keeping in mind that ASSOCIATION there are often many extinct lineages that are not rep- obtained by using a variety of different approaches resented on a phylogeny of extant species; see Nunn, (for reviews, see Kamilar & Cooper, 2013; Münke- 2011). müller et al., 2012; Nunn, 2011). Here we focus on The extent of shared evolutionary history one commonly used metric, Pagel’s lambda (Freckle- between pairs of species in a phylogeny can be rep- ton et al., 2002; Pagel, 1999a). resented as a variance–covariance matrix (see Figure Lambda is a continuous parameter that ranges 10.1B; Cunningham, Omland, & Oakley, 1998; from 0 to 1, with a value of 0 indicating that trait Freckleton, Harvey, & Pagel, 2002). The diagonal covariances are independent of phylogeny and a of the matrix (gray background) represents the vari- valuePSYCHOLOGICAL of 1 indicating that variation among spe- ances, or the total time from the root to the tips of cies approximates expectations from a Brownian the tree (4 million years). The off-diagonals of the motion model of evolution (i.e., a random walk in matrix represent the covariances, or the amount of which trait variance accumulates proportionally to time that pairs of species have shared evolutionary evolutionary time). The lambda parameter scales history since the root of the tree. In this example, the internal branches of a phylogeny by multiply- the covariance between species C and D AMERICANis 3 mil- ing the internal branch lengths (the off-diagonals in lion years, refecting the time between© nodes r and the variance–covariance matrix) by lambda while d, whereas the covariance between species C and retaining the original variances along the diagonal. H is 1, refecting the time between nodes r and b. Thus, when λ = 0, the internal structure of the tree Variance–covariance matrices play an important role is entirely eliminated (all internal branch lengths in many phylogenetic comparativePROOFS methods; hence, are 0, yielding a “star phylogeny”; see Felsenstein, we revisit this concept throughout the chapter. 1985). In contrast, the internal branch lengths (off- A frst step in most comparative analyses is to diagonals in the variance–covariance matrix) remain obtain a phylogeny for the species of interest. For- unchanged when λ = 1. Lambda may take any value tunately, digital phylogenies are widely available between 0 and 1 (and even slightly higher than 1, and can be downloaded from sites such as 10ktrees subject to constraints determined by characteristics (http://10ktrees.fas.harvard.edu; Arnold, Mat- of the tree). The lambda value for a given trait is typically estimated using maximum likelihood to UNCORRECTEDthews, & Nunn, 2010), TreeBase (http://treebase. org; Sanderson, Donoghue, Piel, & Eriksson, 1994), fnd the lambda transformation that makes covari- and the Open Tree of Life (http://blog.opentreeofife. ances between species most likely under a Brownian org). Garamszegi and Gonzalez-Voyer (2014) pro- motion model of evolution. vided an overview of the steps for obtaining a phy- To illustrate the concept of lambda and phylo- logeny for comparative research. genetic signal, Figure 10.2 shows the values of two

3

BK-APA-HCM_V1-160213-Chp10.indd 3 04/07/16 6:13 PM MacLean and Nunn

FIGURE 10.2. Two traits (indicated in gray and black) with differing levels of phylogenetic signal plotted on a phylogeny. The vertical positions of the points are proportional to the trait values. The gray trait has strong phylogenetic signal (closely related species have more similar trait values), and the black trait has weak phylogenetic signal (trait variance is independent of phylogeny). ASSOCIATION continuous traits that were simulated on a phylogeny of lambda for indifference points was ∼1 (see Vol- to have different levels of phylogenetic signal. The ume 2, Chapter 24, this handbook). Similarly, in tests maximum likelihood estimate of lambda for the gray of self-control administered to 36 species, including trait is 0.99, indicating that closely related species primates, carnivores, rodents, and birds, MacLean have highly similar trait values (notice, e.g., that sis- et al. (2014) reported lambda values of 0.76 across ter species have gray bars of approximately the same the entire sample of species (for the average score magnitude). In contrast, the maximum likelihood across cognitive tasks) and 0.86 within primates. estimate of lambda for the black trait is ∼0, indicat- Thus, as with many other traits, phenotypic variation ing that phylogeny is a weak predictor of species-level in cognitionPSYCHOLOGICAL is not independent of phylogeny. variance (hence, one sees no pattern of sister species One common misconception about testing for having black bars of more similar magnitude, com- phylogenetic signal in individual traits is that the pared with other species on the tree). More impor- results of these analyses determine whether phy- tant, these branch-length transformations are not logenetic approaches are needed for subsequent a revised estimate of the actual species divergence analyses. For example, if none of the individual times (the best estimate of which was known beforeAMERICAN traits being studied display phylogenetic signal, analysis). Instead, the transformation estimates© the researchers may be tempted to use these fndings extent to which phenotypic covariance matches the as a justifcation to forego additional phylogenetic expected covariance on the basis of phylogeny and analyses. However, as we show in the next section, a Brownian motion model of evolution. For tests of it is phylogenetic signal not in individual traits but phylogenetic signal in discretePROOFS traits, we direct the in the error variance of a statistical model that is reader to Maddison and Slatkin (1991), Abouheif relevant in comparative analyses of correlated evolu- (1999), and Fritz and Purvis (2010). tion in two or more traits. Of course, a lack of phy- Assessing phylogenetic signal is an important logenetic signal should also lead one to be cautious frst step in comparative analyses because it provides about applying phylogenetic methods to reconstruct information about the extent of nonindependence in ancestral nodes or to assess how traits infuence spe- the data. To date, only a few studies have investigated ciation and extinction rates across the tree. phylogenetic signal in measures of cognition across species.UNCORRECTED Yet the initial fndings have suggested that “HOW” AND “WHY” QUESTIONS: TESTING several cognitive measures are characterized by sub- ADAPTIVE HYPOTHESES stantial phylogenetic signal relative to behavioral or ecological traits. For example, in a comparative analy- Many of the most interesting questions about cogni- sis of primate data from an intertemporal choice task, tive evolution concern when, how, and why particu- Stevens found that the maximum likelihood estimate lar aspects of cognition evolved. “When” questions

4

BK-APA-HCM_V1-160213-Chp10.indd 4 04/07/16 6:13 PM Phylogenetic Approaches for Research in Comparative Cognition

often involve assessing whether a particular charac- in X or Y alone is insuffcient justifcation for ignor- ter state is ancestral in a clade or evolved indepen- ing phylogeny in regression models using these vari- dently. These questions embrace Tinbergen’s (1963) ables (Revell, 2010). question about evolutionary history. “How” ques- Generalized least squares (GLS) is an exten- tions are typically questions about the proximate sion of the general linear model that can accom- mechanisms (Tinbergen, 1963) underlying pheno- modate nonindependence in the data, as well as typic variance. For example, do differences in life non-Gaussian error distributions (see Chapter 8, history strategies, gene expression in the brain (see this volume). Phylogenetic generalized least squares Chapter 22, this volume), or the volume of brain (PGLS) is a regression model that is based on the regions relate to species differences in cognition (see GLS framework (Grafen, 1989; Martins & Hansen, Chapters 12 and 24, this volume)? “Why” questions 1997; Pagel, 1999a). It is similar to GLS in that it are typically questions about the selective pressures generates regression coeffcients, confdence inter- that have favored cognitive evolution. For example, vals, and signifcance estimates for one or more pre- does living in complex societies, relying on spatio- dictors (which can be continuousASSOCIATION or discrete) of a temporally distributed resources, or facing high lev- continuous response variable. PGLS accommodates els of niche competition favor particular cognitive the nonindependence of species-level data by incor- adaptations? porating the variance–covariance matrix (described One of the most powerful approaches for inves- above) into the error term of the model to account tigating how and why questions involves tests of for covariance in model residuals given the phylog- correlated evolution. In brief, the rationale for such eny and the extent of phylogenetic signal in the data tests is that correlations between traits across phy- (Symonds & Blomberg, 2014). logeny suggest that these traits may be function- The fexibility of PGLS lies in its ability to adjust ally linked (Harvey & Purvis, 1991; Nunn, 2011; forPSYCHOLOGICAL statistical nonindependence on the basis of the Nunn & Barton, 2001). However, as we discovered actual error structure of the data (based on estimat- in our review of phylogenetic signal, the noninde- ing lambda or other scaling parameters). This is a pendence of species-level data is an important con- critical difference between PGLS and independent sideration when testing such hypotheses because contrasts, an earlier approach that is used to obtain many traits are correlated at the species level statistically independent comparisons within a phy- through inheritance from a common ancestor,AMERICAN not logeny (for a review, see Felsenstein, 1985; Nunn & independent evolutionary origins. © Barton, 2001). Specifcally, independent con- General linear models provide a unifed statisti- trasts are calculated by generating differences—or cal framework for predicting a continuous depen- contrasts—between pairs of lineages throughout dent measure as a function of an intercept and a the tree. The method assumes that phylogeny accu- linear combination of predictorPROOFS variables weighted rately predicts the model’s error structure and thus by regression coeffcients (see Chapter 8, this vol- standardizes contrasts by dividing them by the sum ume). General linear models assume that errors of their branch lengths to control for heteroske- (model residuals) are normally and independently dasticity. An alternative is to estimate the extent of distributed with common variance (homoscedas- phylogenetic signal in the data and adjust the model ticity). Comparative data frequently violate this accordingly (Symonds & Blomberg, 2014). To do assumption because the residuals of closely related so, PGLS is often combined with estimating branch- UNCORRECTEDspecies tend to be more similar than those of more length transformations (such as Pagel’s lambda) that distantly related species, and thus errors are not accommodate diverse evolutionary models and dif- independently distributed (i.e., they exhibit phylo- ferent levels of phylogenetic nonindependence. As genetic signal). More important, this scenario can in the case for estimating phylogenetic signal in a arise even in cases when neither the predictor nor single trait, maximum likelihood can be used to fnd the response variable has high levels of phylogenetic the branch-length transformation that optimizes the signal, and thus the absence of phylogenetic signal error structure of the residuals in a regression model

5

BK-APA-HCM_V1-160213-Chp10.indd 5 04/07/16 6:13 PM MacLean and Nunn

(Revell, 2010). When lambda is estimated to be 1, association exists between X and Y, as shown in the results of the PGLS model will be identical to the clade-specifc dashed regression lines. Between those from independent contrasts, and when lambda clades, however, grade shifts occur in both X and is estimated to be 0, PGLS results will mirror those Y. When treating species as independent data from a general linear model (or GLS with a Gauss- points (i.e., when lambda is forced to equal 0), GLS ian distribution) that does not include phylogeny. strongly overestimates the relationship between X However, in many cases the maximum likelihood and Y (gray regression line; R2 = .74, p < .001). In estimate of lambda will lie intermediate to these two contrast, PGLS accounts for the nonindependence extremes. in the model and correctly fails to reject the null A common misconception about phylogenetic hypothesis (λ = 0.72, R2 = .05, p = .07). models is that they are more conservative than non- Figure 10.3B shows a different situation in which phylogenetic approaches and thus guard only against PGLS again leads one to a correct interpretation of Type I statistical errors (see Chapter 8, this volume). scaling relationships. In this case, there is a grade However, by meeting the assumptions of the under- shift in X but not in Y, which whenASSOCIATION analyzed by lying statistical model, phylogenetic approaches not GLS (gray regression line) obscures the strong rela- only reduce false positives by correcting for pseudo- tionships between X and Y within each of the three replication but also have increased power to detect clades (dashed regression lines). In this case, GLS correlated evolution in cases in which a general fails to detect the correlation between X and Y (R2 = linear model does not do so (Type II errors). Both of .05, p = .10), whereas PGLS has increased power to these scenarios are illustrated in hypothetical exam- do so (λ = 0.84, R2 = .60, p < .001). ples exploring the relationship between an ecological As noted previously, PGLS can accommodate predictor and a cognitive dependent measure in both continuous and discrete predictors of a contin- Figure 10.3. Figure 10.3A depicts the classic prob- uous responsePSYCHOLOGICAL variable, but it is inappropriate to use lem of pseudoreplication (see Felsenstein, 1985). PGLS for a discrete response variable (i.e., logistic Data from three clades (monophyletic groups) are regression). For a review of methods appropriate for shown, with clade membership indicated by the categorical response variables, we direct the reader shading of points. Within each clade, no signifcant to Ives and Garland (2014). AMERICAN (A) © (B)

PROOFS Cognitive Measure Cognitive Measure

Ecological Predictor Ecological Predictor UNCORRECTED FIGURE 10.3. Examples of scenarios in which phylogenetic generalized least squares avoids Type I (A) and Type II (B) statistical errors. Clade mem- bership is represented by the color of points. The solid gray regression line is from a nonphylogenetic model that fails to account for the nonindependence of species-level data. The dashed black lines show the clade-specifc associa- tions between the ecological predictor and cognitive dependent measure. See text for details.

6

BK-APA-HCM_V1-160213-Chp10.indd 6 04/07/16 6:13 PM Phylogenetic Approaches for Research in Comparative Cognition

PGLS has been incorporated in several recent of humans and great apes have a theory of mind? studies of comparative cognition. For example, When, and how many times, has social learning using PGLS with a sample of 13 primate species, evolved? Do vertebrate species share a primitive Stevens (2014) identifed links between a range of number sense? allometric variables, including body mass, brain The answers to these types of questions rely volume, life span, and home range size, and spe- on the ability to make inferences about the past cies differences in a delayed reward task. In another using information from the present and thus cap- study, MacLean et al. (2014) found that absolute ture the when questions identifed above. Several brain volume and dietary breadth, but not frugivory comparative methods are designed for ancestral or species-typical social group size, predicted differ- state reconstructions. Here, we focus on maximum ences on two self-control tasks among 36 species of likelihood approaches for inferences about discrete mammals and birds. Interestingly, when a smaller traits, although similar methods can be used with subset of species was tested on both a social cogni- continuous variables (for review, see MacLean et al., tive measure and one of these self-control tasks, 2012). Maximum likelihood methodsASSOCIATION differ from MacLean et al. (2013) found that social group size parsimony reconstructions in a number of impor- was a signifcant predictor of species differences tant ways (Pagel, 1999b). Reconstructions based on the social but not the self-control task. Last, in on parsimony attempt to minimize the number of a recent comparative study of 15 primate species transitions on the tree, and they ignore information (including humans), Burkart et al. (2014) found about branch lengths (i.e., changes are assumed to that extensive allomaternal care was the best predic- be equally likely to occur along the shortest and the tor of proactive prosociality, implicating cooperative longest branch on the tree). Maximum parsimony breeding as a possible selective pressure for human methods also do not provide easily interpretable hypercooperation (see Chapter 13, this volume). statisticalPSYCHOLOGICAL support measures for reconstructed states We return to this example below to assess whether (Nunn, 2011). More important, parsimony lacks an humans show substantially more prosociality than underlying evolutionary model and is being driven other primates given their level of allomaternal care to extinction by model-based approaches, such as (see Phylogenetic Prediction section). those based on maximum likelihood. By using a Thus, PGLS has allowed researchers to test a specifc evolutionary model, maximum likelihood wide range of hypotheses about cognitiveAMERICAN evolution. approaches integrate information about branch Unlike studies using proxy variables for© cognition, lengths (the amount of time that lineages have such as brain size, a combination of experimental evolved independently of other taxa in the tree) and phylogenetic methods has led to unique insights and rates of evolutionary change, and they can pro- about the selective pressures that have favored cog- vide estimates of statistical support for each recon- nitive evolution in speciPROOFSfc cognitive domains. Col- structed node in the tree. These models can also lectively, these types of studies highlight an exciting accommodate specifc types of trait evolution, such new direction for research in comparative cognition, as asymmetric transition rates in which the prob- one that has potential to lead to major insights about ability of moving from state A to state B differs from long-standing questions regarding how and why that of moving from state B to state A. cognition evolves. To illustrate a maximum likelihood reconstruc- tion using discrete variables, we present an example using data on social learning, extractive foraging, UNCORRECTED“WHEN” QUESTIONS: RECONSTRUCTING and tool use taken from Reader, Jager, and Laland THE COGNITIVE PAST (2011). This dataset includes counts of published Comparative psychologists study the behavior and reports for each of these categories in a large sample cognition of extant species, yet many questions in of primates, along with the total number of citations comparative psychology involve inferences about for each species at the time data were compiled. the past. For example, did the last common ancestor To create discrete traits based on these data, we

7

BK-APA-HCM_V1-160213-Chp10.indd 7 04/07/16 6:13 PM MacLean and Nunn

frst restricted the dataset to species with at least 20 member of the genus (each of these coded as yes citations in the Zoological Record database, effec- [1] or no [0]). Ancestral states were then estimated tively limiting the data to a subset of reasonably using maximum likelihood in a model with equal well-studied species. Species-level data were then transition rates between the character states for each collapsed to a genus-level summary (44 genera) to variable, meaning that the rates of state yes transi- allow a concise example for illustrative purposes. tioning to state no were equivalent to rates of state Finally, within each genus we created three discrete no transitioning to state yes. variables refecting whether social learning, extrac- The results of this analysis are plotted on the genus- tive foraging, and tool use had been observed in any level tree in Figure 10.4. The tips of the tree depict the e forag tool use social learning ext. ASSOCIATION Oolemur Nycticebus Daubentonia Varecia Lemur Hapalemur Eulemur Propithecus Microcebus Pithecia Chiropotes tool use social learning ext. forage ext. Saimiri Cebus Saguinus Leontopithecus PSYCHOLOGICALCallithrix Callimico Aotus Lagothrix Brachyteles Ateles Alouatta Pongo Pan Gorilla AMERICAN Nomascus Symphalangus © Hylobates Bunopithecus Trchypithecus Semnopithecus Rhinopithecus Pygathrix Piliocolobus PROOFS Colobus Macaca Papio Theropithecus Lophocebus Mandrillus Cercocebus Chlorocebus Erythrocebus Cercopithecus

FIGURE 10.4. Ancestral state reconstruction of tool use, extractive foraging, UNCORRECTEDand social learning modeled as discrete traits in primate genera. Squares at the tips of the phylogeny represent the character states for extant species (shaded = trait present, unshaded = trait absent). The circles at the internal nodes represent the scaled likelihood that a trait was present at that node (proportion shaded = likelihood a trait was present). The leftmost circles and squares correspond to tool use. The center circles and squares correspond to extractive foraging. The rightmost circles and squares correspond to social learning. See text for details.

8

BK-APA-HCM_V1-160213-Chp10.indd 8 04/07/16 6:13 PM Phylogenetic Approaches for Research in Comparative Cognition

character states for extant genera using shaded (trait that rely on convergence to infer adaptation. How- present) or unshaded (trait absent) squares (left, tool ever, comparative methods can be used to determine use; center, extractive foraging; right, social learning). whether a particular species deviates from what one The circles at the internal nodes of the phylogeny indi- expects given the phylogeny and given the relation- cate the scaled likelihood (range = 0–1) that each of ships between a set of predictor and response vari- the three traits was present at that node. ables measured in other taxa (Nunn & Zhu, 2014; This analysis indicates that the last common Organ, Nunn, Machanda, & Wrangham, 2011). For ancestor of the extant great apes used tools (see Vol- example, human brains may have more neurons ume 2, Chapter 30, this handbook), practiced extrac- than any other primate species, but this trait may tive foraging, and learned socially (see Volume 2, not be extraordinary given knowledge of human Chapter 19, this handbook), effectively answering a brain volumes and the general cellular scaling rules when question about these behaviors. At the root of of primate brains (Herculano-Houzel, 2009, 2012). the phylogeny, much more uncertainty exists about Many long-standing questions in comparative the character states of all three variables; this is a psychology relate to whether aspectsASSOCIATION of cognition common circumstance, which often requires that are qualitatively different in humans and nonhu- additional outgroups be added to the tree to bet- man animals (Penn, Holyoak, & Povinelli, 2008; ter resolve the state of the character at the deeper Shettleworth, 2012). Often, these questions are nodes. Visualizing the trends across the phylogeny, investigated by comparing the performance of one sees that tool use is estimated to be a recently humans and nonhuman animals on a set of cogni- evolved trait (compared with social learning and tive measures with the aim of identifying areas of extractive foraging), and a trait that appears only similarity and difference (e.g., Herrmann, Call, after both social learning and extractive foraging are Hernàndez-Lloreda, Hare, & Tomasello, 2007; well established. Thus, one plausible explanation is InouePSYCHOLOGICAL & Matsuzawa, 2007). However, even in cases that tool use is adaptive primarily in species already in which humans do differ substantially from other feeding on diffcult-to-access or embedded foods species, these differences may be predictable within and that social learning may be required for tool use a phylogenetic framework and, in this sense, con- to become widespread (i.e., it is acquired through sistent with broader evolutionary patterns. Methods social learning). This hypothesis could be evalu- for phylogenetic prediction allow statistical infer- ated using a test for correlated evolution AMERICANof discrete ences about the extent to which an apparent outlier traits, in which the state of one variable© infuences is truly remarkable given phylogeny and a set of rel- the probability of change in another (Pagel, 1994). evant predictor variables. However, a recent article urged caution in using this As an example of phylogenetic prediction, we and some other methods of correlated evolution for consider whether humans are unique in their pro- discrete characters (MaddisonPROOFS & FitzJohn, 2015). social tendencies (see Chapters 13 and 44, this vol- Historically, estimating ancestral states has not ume). Our analysis here builds on recent research been a common approach in comparative psychol- by Burkart et al. (2014), who focused on accounting ogy. However, we expect that with growing com- for variation in data from prosociality experiments parative databases on animal cognition, these types in 15 primates species. Across species, the extent of of analysis will become increasingly common. For allomaternal care was the best predictor of proac- empirical examples of ancestral state estimation for tive prosociality. Humans exhibited very high levels UNCORRECTEDcontinuous cognitive traits, we direct the reader to of proactive prosociality, and they deviated from MacLean et al. (2014) and Reader et al. (2011). the nonhuman primate regression line by less than 1 standard deviation, although this latter analysis does not account for phylogeny in assessing whether PHYLOGENETIC PREDICTION this residual is extreme. Here, we use more power- Evolutionary singularities—or uniquely derived ful methods to assess whether humans are different traits—present a challenge for comparative methods from other primates, accounting for variation in

9

BK-APA-HCM_V1-160213-Chp10.indd 9 04/07/16 6:13 PM MacLean and Nunn

allomaternal care and phylogenetic placement of distribution. Thus, in addition to generating a point humans. We also use this opportunity to introduce estimate of particular parameters (e.g., the mean Bayesian methods, which are becoming increasingly value across saved iterations), uncertainty about common in phylogenetic methodology. these values is refected in the distribution of values A number of approaches are available for assess- that are recorded (“sampled”). In a Bayesian frame- ing whether a species (or set of species) exhibits work, this is referred to as a posterior (as opposed to an exceptionally large (or small) state of a continu- a prior) probability distribution. ously varying trait, such as body mass, height, or Using MCMC in a phylogenetic regression, even proactive prosocial tendencies. Focusing on one can sample the posterior probability distribu- humans, we could investigate whether the lineage tion of regression coeffcients, the intercept, and leading to humans shows a higher rate of evolution lambda that are central to prediction in the PGLS in prosocial behavior compared with other lineages model while also sampling from across a set of trees on the phylogeny. This can be achieved using inde- that refect phylogenetic uncertainty (Nunn, 2011; pendent contrasts (McPeek, 1995) or by ftting Pagel & Lutzoni, 2002). The analysisASSOCIATION takes into models with two different rates for different lineages account all these sources of uncertainty in coef- and assessing whether that two-rate model improves fcients, phylogeny, and scaling parameters; the on a model that fts a single rate (O’Meara, Ané, resulting set of parameters is then used to generate Sanderson, & Wainwright, 2006; Revell, 2008). a posterior probability prediction of, for example, Other, more sophisticated variants on these methods the state of prosocial behavior in humans, predicted are now available (Revell, Mahler, Peres-Neto, & on the basis of human levels of allomaternal care Redelings, 2012). One could also examine residuals and humans’ relationship to other primates. This from a regression of prosocial tendencies on another posterior probability distribution is a natural way trait, such as allomaternal care, with the expectation to incorporatePSYCHOLOGICAL uncertainty in our prediction. If the that humans would have an exceptionally high posi- true human value of prosocial behavior falls outside tive residual (i.e., they are more prosocial than one the 95% credible interval of the posterior probabil- might expect given the degree of allomaternal care ity distribution for predicted prosocial behavior, observed in humans). To incorporate phylogeny, we would say that human prosocial behavior is this latter approach can be implemented with inde- exceptional relative to other primates, meaning that pendent contrasts or PGLS (Garland & Ives, AMERICAN2000). humans are an evolutionary outlier (Nunn, 2011; Here, we use a new approach to phylogenetic© Nunn & Zhu, 2014; Organ et al., 2011). prediction that involves one or more predictor We used the program BayesModelS to run the variables in a PGLS framework but runs the analy- MCMC analysis (Nunn & Zhu, 2014). The pro- sis using a Bayesian approach implemented with gram takes input data on prosocial behavior and Markov chain Monte Carlo, PROOFSor MCMC (Nunn & allomaternal care, which we obtained from Burkart Zhu, 2014). Briefy, at each iteration, the procedure et al. (2014), and a posterior probability distribu- evaluates a new candidate set of parameter values tion of trees, which we obtained from Version 3 of that are accepted (replacing the current values in the 10kTrees (Arnold et al., 2010). The user must also chain) or rejected (maintaining the current values decide how large of a posterior probability sample in the chain). In general, the algorithm is designed to obtain (i.e., how many samples to save) and to select values that make the data more likely, but how often to sample from the chain of parameters lessUNCORRECTED likely solutions may be accepted probabilisti- that is produced. Sampling can too often lead to a cally in relation to the likelihood (i.e., parameter high correlation among the estimates that are used sets that make the data less likely are less likely to be in the posterior probability distribution, whereas accepted). A record of selected parameter values is sampling too little can result in a longer run time. recorded at predefned intervals (e.g., every 100 iter- The user must also ensure that the sample has ations). The frequency with which particular values reached a steady state of estimates (i.e., that it is are sampled ultimately approximates a probability a post–“burn-in” set of parameters during which

10

BK-APA-HCM_V1-160213-Chp10.indd 10 04/07/16 6:13 PM Phylogenetic Approaches for Research in Comparative Cognition

likelihoods increase rapidly across iterations as the allomaternal care to prosocial behavior is shown in model parameters are initially tuned to the data). On Figure 10.5D. It can be seen that the vast majority the basis of initial runs of the model, we decided to of the coeffcients are greater than zero, which is sample 8,000 estimates of the parameters, with 50 strongly indicative of a positive association. There- generations separating each saved set of parameters fore, the model using nonhuman primates suggests and an initial burn-in of 100 generations (in which that allomaternal care is one reason why proactive no samples were retained). prosociality may have increased in certain lineages We also estimated both lambda and kappa, where (Burkart et al., 2014). Finally, we found support for the latter parameter scales branches by raising them both lambda and kappa transformations, with most to the power kappa. We allowed BayesModelS to use of the chain favoring estimation of lambda rather an MCMC procedure to decide whether to estimate than kappa, and a relatively fat distribution of lambda or kappa, with the proportion of samples kappa (see Figures 10.5E and 10.5F). estimating one or the other giving a measure of rela- With posterior probability distributions of coef- tive support for one transformation over the other. fcients, kappa or lambda, andASSOCIATION intercepts, we then Likewise, we allowed the model to estimate the predicted the value of prosocial behavior in humans, regression coeffcient for the effect of allomaternal using our phylogenetic placement and value of care on prosocial behavior or to set it to zero; the allomaternal care (3.56, untransformed) for each proportion of time that the coeffcient was included of the 8,000 stored coeffcients and parameters. in the model and estimated, along with the cred- This analysis produced 8,000 predictions, sum- ible interval on the estimate, gives a sense of how marized by the posterior probability distribution strongly prosocial behavior covaries positively with for logit-transformed data shown in Figure 10.5G, allomaternal care. Finally, we transformed the data. which can be compared with the distribution of We used the logit transformation to rescale the pro- logit-transformedPSYCHOLOGICAL prosocial behavior for all pri- social behavior from a truncated distribution in per- mates (see Figure 10.5H). As can be seen, humans centage or proportional terms to negative infnity to are not exceptionally different from other primates positive infnity (for extremes of 0 and 1 proportion (shown as the black vertical line in Figure 10.5G), of successes, respectively). For two species (Macaca as might be expected given the small residual noted silenus and Varecia variegate), values of zero for pro- above. Thus, although the value of prosocial behav- social behavior were set to 0.001 before applyingAMERICAN the ior in humans is slightly higher than expected, it is logit to avoid negative infnity values.© For allomater- clearly not an outlier given human levels of alloma-

nal care, we used a log10 transformation. ternal care and humans’ placement in the primate The frst stage of our analysis involves estimating phylogeny. the association between prosocial behavior and allo- maternal care. The MCMCPROOFS model reached stationar- CONCLUDING REMARKS ity (relatively consistent likelihood estimates after the initial burn-in) and showed a good distribution The comparative methods reviewed in this chapter of likelihoods (see Figure 10.5A), with most of the provide an essential tool kit for investigating a wide chain sampling parameters that made the data more range of questions about when, why, and how cogni- likely but also sampling some less likely parameter tion evolves. Coupled with high-quality datasets on combinations (less often). We found strong evi- species differences in cognition, this approach has UNCORRECTEDdence that allomaternal care should be included in the potential to catalyze an exciting revolution in the statistical model with prosocial behavior as a comparative psychology. More important, the success response variable, with 98.2% of the models includ- of this endeavor will require datasets covering larger ing allomaternal care, and a generally higher likeli- comparative samples than are currently common in hood in models that included allomaternal care (see this feld. For example, several of the methods we Figures 10.5B and 10.5C). The posterior probability reviewed are known to be highly sensitive to sample distribution of the regression coeffcient relating size, having limited statistical power in samples

11

BK-APA-HCM_V1-160213-Chp10.indd 11 04/07/16 6:13 PM MacLean and Nunn

(A) (E) –20 –40 –60 –80 300

Likelihood –100 0 –120 Frequency 0 2000 4000 6000 8000 0.0 0.2 0.4 0.6 0.8 1.0 Iteration Kappa (n = 2210)

(B) (F) With Allomaternal Care

1500 500

Frequency 0 0 Frequency –80 –70 –60 –50 –40 –30 –20 –10 0.0 0.2 0.4 0.6 0.8 1.0 Log likelihood Lambda (n = 5790) (C) (G) Without Allomaternal Care 1200 ASSOCIATION

40 800 400 Frequency

Frequency 0 0 –80 –70 –60 –50 –40 –30 –20 –10 –2 0 2 4 Predicted Proactive Prosociality (Human) (D) Log Likelihood (H) 3.0

1000 2.0 500 1.0 PSYCHOLOGICAL Frequency 0 Frequency 0 –5 0510 15 20 –2 0 2 4 Regression Coefficient Observed Proactive Prosociality (Primates)

FIGURE 10.5. Phylogenetic prediction of proactive prosociality in humans. A: Distribution of likelihoods basedAMERICAN on the Markov chain Monte Carlo model. B: Likelihoods for models including© allomaternal care. C: Likelihoods for models excluding allomaternal care. D: Posterior probability distribution of regres- sion coeffcients relating allomaternal care to proactive prosociality. E and F: Distributions of kappa and lambda across model fts. G and H: Posterior probabil- ity distribution for human proactive prosociality compared with observed varia- tion in nonhuman primates. The vertical black line (G) indicates the observed value for humans.PROOFS See text for details.

smaller than 20 to 30 species (Kamilar & Cooper, advanced topics (see Table 10.1). Although compara- 2013; Münkemüller et al., 2012). Thus, phylogenetic tive methods can be implemented in many different approaches for the study of comparative psychology software packages, which range in the extent of their will require not only special statistical methods but fexibility, the amount of coding required by the user, alsoUNCORRECTED datasets that are appropriate for the implementa- and documentation, the vast majority of compara- tion of these techniques. In this regard, we expect tive methods (including all of those presented in this that large-scale collaboration will also be an essential chapter) are available in R (http://www.r-project. ingredient for success (MacLean et al., 2012, 2014). org), a free and open-source language for statistical For readers interested in learning more about computing. Table 10.1 provides a list of useful pack- comparative methods, a wide variety of freely avail- ages for phylogenetic analysis using R, and tutorials able resources now exist that cover both basic and including example code for the methods presented in

12

BK-APA-HCM_V1-160213-Chp10.indd 12 04/07/16 6:13 PM Phylogenetic Approaches for Research in Comparative Cognition

TABLE 10.1

Online Resources for Phylogenetic Comparative Methods

Resource URL Reference

Websites

AnthroTree http://www.anthrotree.info Nunn (2011) 10kTrees Project http://10ktrees.fas.harvard.edu Arnold, Matthews, and Nunn (2010) TreeBASE http://treebase.org Sanderson, Donoghue, Piel, and Eriksson (1994) Open Tree of Life http://tree.opentreeoflife.org Hinchliff et al. (2014) Programming language R Project for Statistical Computing http://www.r-project.org R Core Team (2014) R packages ASSOCIATION APE http://cran.r-project.org/package=ape Paradis, Claude, and Strimmer (2004) CAPER http://cran.r-project.org/package=caper Orme et al. (2011) Phytools http://cran.r-project.org/package=phytools Revell (2011) Geiger http://cran.r-project.org/package=geiger Harmon et al. (2009)

this chapter can be accessed online at the AnthroTree Blomberg, S. P., Garland, T., Jr., & Ives, A. R. (2003). website (http://www.anthrotree.info). Testing for phylogenetic signal in comparative data: Behavioral traits are more labile. Evolution: In conclusion, phylogenetic comparative meth- PSYCHOLOGICALInternational Journal of Organic Evolution, 57, 717–745. ods will play a key role in unraveling the natural http://dx.doi.org/10.1111/j.0014-3820.2003.tb00285.x history of cognition and inferring the historical Burkart, J. M., Allon, O., Amici, F., Fichtel, C., processes that have shaped the minds of the species Finkenwirth, C., Heschl, A., . . . van Schaik, C. P. we study today. We hope the topics reviewed in this (2014). The evolutionary origin of human hyper- chapter will stimulate further interest in phyloge- cooperation. Nature Communications, 5, 4747. http:// dx.doi.org/10.1038/ncomms5747 netic approaches to the study of comparativeAMERICAN psy- chology, and we look forward to future© discoveries Cunningham, C. W., Omland, K. E., & Oakley, T. H. (1998). Reconstructing ancestral character states: A emerging from this approach. critical reappraisal. Trends in Ecology and Evolution, 13, 361–366. http://dx.doi.org/10.1016/S0169- References 5347(98)01382-2 PROOFS Abouheif, E. (1999). A method for testing the assumption Felsenstein, J. (1985). Phylogenies and the comparative of phylogenetic independence in comparative data. method. American Naturalist, 125, 1–15. http:// Evolutionary Ecology Research, 1, 895–909. dx.doi.org/10.1086/284325 Amici, F., Aureli, F., & Call, J. (2008). Fission-fusion Freckleton, R. P., Harvey, P. H., & Pagel, M. (2002). dynamics, behavioral fexibility, and inhibitory Phylogenetic analysis and comparative data: A test control in primates. Current Biology, 18, 1415–1419. and review of evidence. American Naturalist, 160, http://dx.doi.org/10.1016/j.cub.2008.08.020 712–726. http://dx.doi.org/10.1086/343873 Arnold, C., Matthews, L. J., & Nunn, C. L. (2010). The Fritz, S. A., & Purvis, A. (2010). Selectivity in UNCORRECTED10k trees website: A new online resource for primate mammalian extinction risk and threat types: A new phylogeny. Evolutionary Anthropology, 19, 114–118. measure of phylogenetic signal strength in binary http://dx.doi.org/10.1002/evan.20251 traits. Conservation Biology, 24, 1042–1051. http:// dx.doi.org/10.1111/j.1523-1739.2010.01455.x Blomberg, S. P., & Garland, T., Jr. (2002). Tempo and mode in evolution: Phylogenetic inertia, adaptation Garamszegi, L. Z. (Ed.). (2014). Modern phylogenetic and comparative methods. Journal of Evolutionary comparative methods and their application in Biology, 15, 899–910. http://dx.doi.org/10.1046/ evolutionary biology. http://dx.doi.org/10.1007/978-3- j.1420-9101.2002.00472.x 662-43550-2

13

BK-APA-HCM_V1-160213-Chp10.indd 13 04/07/16 6:13 PM MacLean and Nunn

Garamszegi, L. Z., & Gonzalez-Voyer, A. (2014). methods and their application in evolutionary biology Working with the tree of life in comparative studies: (pp. 231–261). http://dx.doi.org/10.1007/978-3-662- How to build and tailor phylogenies to interspecifc 43550-2 datasets. In L. Z. Garamszegi (Ed.), Modern Kamilar, J. M., & Cooper, N. (2013). Phylogenetic signal phylogenetic comparative methods and their application in primate behaviour, ecology and life history. (pp. 19–48). http://dx.doi. in evolutionary biology Philosophical Transactions of the Royal Society of org/10.1007/978-3-662-43550-2 London. Series B, Biological Sciences, 368, 20120341. Garland, T., Jr., Bennett, A. F., & Rezende, E. L. (2005). http://dx.doi.org/10.1098/rstb.2012.0341 Phylogenetic approaches in comparative . MacLean, E. L., Hare, B., Nunn, C. L., Addessi, E., Journal of Experimental Biology, 208, 3015–3035. Amici, F., Anderson, R. C., . . . Zhao, Y. (2014). The http://dx.doi.org/10.1242/jeb.01745 evolution of self-control. Proceedings of the National Garland, T., Jr., & Ives, A. R. (2000). Using the past Academy of Sciences, USA, 111, E2140–E2148. http:// to predict the present: Confdence intervals for dx.doi.org/10.1073/pnas.1323533111 regression equations in phylogenetic comparative MacLean, E. L., Matthews, L. J., Hare, B. A., Nunn, C. L., methods. American Naturalist, 155, 346–364. http:// Anderson, R. C., Aureli, F., . . . Wobber, V. (2012). dx.doi.org/10.1086/303327 How does cognition evolve? Phylogenetic comparative Grafen, A. (1989). The phylogenetic regression. psychology. Animal Cognition, 15,ASSOCIATION 223–238. http:// Philosophical Transactions of the Royal Society of dx.doi.org/10.1007/s10071-011-0448-8 London. Series B, Biological Sciences, 326, 119–157. MacLean, E. L., Sandel, A. A., Bray, J., Oldenkamp, R. http://dx.doi.org/10.1098/rstb.1989.0106 E., Reddy, R. B., & Hare, B. A. (2013). Group size Harmon, L., Weir, J., Brock, C., Glor, R., Challenger, W., & predicts social but not nonsocial cognition in lemurs. Hunt, G. (2009). Geiger: Analysis of evolutionary PLOS ONE, 8, e66359. http://dx.doi.org/10.1371/ diversifcation [Computer software]. Retrieved from journal.pone.0066359 http://cran.r-project.org/package=geiger Maddison, W. P., & FitzJohn, R. G. (2015). The Harvey, P. H., & Purvis, A. (1991). Comparative methods unsolved challenge to phylogenetic correlation for explaining adaptations. Nature, 351, 619–624. tests for categorical characters. Systematic Biology, http://dx.doi.org/10.1038/351619a0 64,PSYCHOLOGICAL 127–136. http://dx.doi.org/10.1093/sysbio/ Herculano-Houzel, S. (2009). The human brain in syu070 numbers: A linearly scaled-up primate brain. Maddison, W. P., & Slatkin, M. (1991). Null models for Frontiers in Human Neuroscience, 3, 31. http://dx.doi. the number of evolutionary steps in a character on org/10.3389/neuro.09.031.2009 a phylogenetic tree. Evolution: International Journal Herculano-Houzel, S. (2012). The remarkable, yet not of Organic Evolution, 45, 1184–1197. http://dx.doi. extraordinary, human brain as a scaled-up primateAMERICAN org/10.2307/2409726 brain and its associated cost. Proceedings of ©the National Martins, E. P., & Garland, T. (1991). Phylogenetic Academy of Sciences, USA, 109(Suppl. 1), 10661–10668. analyses of the correlated evolution of continuous http://dx.doi.org/10.1073/pnas.1201895109 characters: A simulation study. Evolution: Herrmann, E., Call, J., Hernàndez-Lloreda, M. V., Hare, International Journal of Organic Evolution, 45, B., & Tomasello, M. (2007). Humans have evolved 534–557. http://dx.doi.org/10.2307/2409910 specialized skills of social cognition:PROOFS The cultural Martins, E. P., & Hansen, T. F. (1997). Phylogenies intelligence hypothesis. Science, 317, 1360–1366. and the comparative method: A general approach http://dx.doi.org/10.1126/science.1146282 to incorporating phylogenetic information into the Hinchliff, C. E., Smith, S. A., Allman, J. F., Burleigh, analysis of interspecifc data. American Naturalist, J. G., Chaudhary, R., Coghill, L. M., . . . Gude, K. 149, 646–667. http://dx.doi.org/10.1086/286013 (2015). Synthesis of phylogeny and taxonomy into a McPeek, M. A. (1995). Testing hypotheses about comprehensive tree of life. Proceedings of the National evolutionary change on single branches of Academy of Sciences, USA, 112, 12764–12769. http:// a phylogeny using evolutionary contrasts. dx.doi.org/10.1073/pnas.1423041112 American Naturalist, 145, 686–703. http://dx.doi. Inoue,UNCORRECTED S., & Matsuzawa, T. (2007). Working memory org/10.1086/285763 of numerals in chimpanzees. Current Biology, Münkemüller, T., Lavergne, S., Bzeznik, B., Dray, 17, R1004–R1005. http://dx.doi.org/10.1016/j. cub.2007.10.027 S., Jombart, T., Schiffers, K., & Thuiller, W. (2012). How to measure and test phylogenetic Ives, A. R., & Garland, T. J. (2014). Phylogenetic signal. Methods in Ecology and Evolution, 3, regression for binary dependent variables. In L. Z. 743–756. http://dx.doi.org/10.1111/j.2041- Garamszegi (Ed.), Modern phylogenetic comparative 210X.2012.00196.x

14

BK-APA-HCM_V1-160213-Chp10.indd 14 04/07/16 6:13 PM Phylogenetic Approaches for Research in Comparative Cognition

Nunn, C. L. (2011). The comparative approach in Penn, D. C., Holyoak, K. J., & Povinelli, D. J. (2008). evolutionary anthropology and biology. http://dx.doi. Darwin’s mistake: Explaining the discontinuity org/10.7208/chicago/9780226090009.001.0001 between human and nonhuman minds. Behavioral Nunn, C. L., & Barton, R. A. (2001). Comparative and Brain Sciences, 31, 109–130. http://dx.doi. methods for studying primate adaptation and org/10.1017/S0140525X08003543 . Evolutionary Anthropology, 10, 81–98. R Core Team. (2014). R: A language and environment for http://dx.doi.org/10.1002/evan.1019 statistical computing. Vienna, Austria: R Foundation Nunn, C. L., & Zhu, L. (2014). Phylogenetic prediction for Statistical Computing. to identify “evolutionary singularities.”. In L. Z. Reader, S. M., Hager, Y., & Laland, K. N. (2011). Garamszegi (Ed.), Modern phylogenetic comparative The evolution of primate general and cultural methods and their application in evolutionary biology intelligence. Philosophical Transactions of the Royal (pp. 481–514). http://dx.doi.org/10.1007/978-3-662- Society of London. Series B, Biological Sciences, 366, 43550-2 1017–1027. O’Meara, B. C., Ané, C., Sanderson, M. J., & Revell, L. J. (2008). On the analysis of evolutionary Wainwright, P. C. (2006). Testing for different change along single branches in a phylogeny. rates of continuous trait evolution using American Naturalist, 172, 140–147. http://dx.doi. likelihood. Evolution: International Journal of org/10.1086/588078 ASSOCIATION Organic Evolution, 60, 922–933. http://dx.doi. Revell, L. J. (2010). Phylogenetic signal and linear org/10.1111/j.0014-3820.2006.tb01171.x regression on species data. Methods in Ecology and Organ, C., Nunn, C. L., Machanda, Z., & Wrangham, Evolution, 1, 319–329. http://dx.doi.org/10.1111/ R. W. (2011). Phylogenetic rate shifts in feeding j.2041-210X.2010.00044.x time during the evolution of Homo. Proceedings Revell, L. J. (2011). Phytools: Phylogenetic tools for of the National Academy of Sciences, USA, 108, comparative biology (and other things) (Version 14555–14559. http://dx.doi.org/10.1073/ 0.0-8) [Computer software]. Retrieved from http:// pnas.1107806108 cran.r-project.org/package=phytools Orme, D., Freckleton, R., Thomas, G., Petzoldt, T., Fritz, Revell,PSYCHOLOGICAL L. J., Mahler, D. L., Peres-Neto, P. R., & Redelings, S., & Isaac, N. (2011). CAPER: Comparative analyses B. D. (2012). A new phylogenetic method for of and evolution in R (Version 0.4/ identifying exceptional phenotypic diversifcation. r64) [Computer software]. Retrieved from http:// Evolution: International Journal of Organic Evolution, cran.r-project.org/package=caper 66, 135–146. http://dx.doi.org/10.1111/j.1558- Pagel, M. (1994). Detecting correlated evolution on 5646.2011.01435.x phylogenies: A general method for the comparative Rezende, E. L., & Diniz-Filho, J. A. F. (2012). analysis of discrete characters. ProceedingsAMERICAN of Phylogenetic analyses: Comparing species to the Royal Society of London. Series B, Biological infer adaptations and physiological mechanisms. Sciences, 255, 37–45. http://dx.doi.org/10.1098/© Comprehensive Physiology, 2, 639–674. rspb.1994.0006 Sanderson, M., Donoghue, M., Piel, W., & Eriksson, Pagel, M. (1999a). Inferring the historical patterns of T. (1994). Treebase: A prototype database of biological evolution. Nature, 401, 877–884. http:// phylogenetic analyses and an interactive tool for dx.doi.org/10.1038/44766PROOFS browsing the phylogeny of life. American Journal of Pagel, M. (1999b). The maximum likelihood Botany, 81, 183. approach to reconstructing ancestral character Shettleworth, S. J. (2012). Modularity, comparative states of discrete characters on phylogenies. cognition and human uniqueness. Philosophical Systematic Biology, 48, 612–622. http://dx.doi. Transactions of the Royal Society of London. Series B, org/10.1080/106351599260184 Biological Sciences, 367, 2794–2802. http://dx.doi. org/10.1098/rstb.2012.0211 Pagel, M., & Lutzoni, F. (2002). Accounting for phylogenetic uncertainty in comparative studies Shultz, S., & Dunbar, R. I. M. (2010). Species differences of evolution and adaptation. In M. Lässig & A. in executive function correlate with hippocampus UNCORRECTEDValleriani (Eds.), Biological evolution and statistical volume and neocortex ratio across nonhuman physics (pp. 148–161). http://dx.doi.org/10.1007/3- primates. Journal of Comparative Psychology, 124, 540-45692-9_8 252–260. http://dx.doi.org/10.1037/a0018894 Paradis, E., Claude, J., & Strimmer, K. (2004). APE: Stevens, J. R. (2014). Evolutionary pressures on primate Analyses of phylogenetics and evolution in R intertemporal choice. Proceedings of the Royal language. Bioinformatics, 20, 289–290. http://dx.doi. Society of London. Series B, Biological Sciences, 281, org/10.1093/bioinformatics/btg412 20140499.

15

BK-APA-HCM_V1-160213-Chp10.indd 15 04/07/16 6:13 PM MacLean and Nunn

Symonds, M. R., & Blomberg, S. P. (2014). A primer Tinbergen, N. (1963). On aims and methods of on phylogenetic generalised least squares. In L. Z. ethology. Zeitschrift für Tierpsychologie, 20, Garamszegi (Ed.), Modern phylogenetic comparative 410–433. http://dx.doi.org/10.1111/j.1439-0310. methods and their application in evolutionary biology 1963.tb01161.x (pp. 105–130). http://dx.doi.org/10.1007/978-3-662- 43550-2

ASSOCIATION

PSYCHOLOGICAL

AMERICAN ©

PROOFS

UNCORRECTED

16

BK-APA-HCM_V1-160213-Chp10.indd 16 04/07/16 6:13 PM