19
University of Groningen Inferring Diversification Rate Variation From Phylogenies With Fossils Mitchell, Jonathan S; Etienne, Rampal S; Rabosky, Daniel L Published in: Systematic biology DOI: 10.1093/sysbio/syy035 IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2019 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Mitchell, J. S., Etienne, R. S., & Rabosky, D. L. (2019). Inferring Diversification Rate Variation From Phylogenies With Fossils. Systematic biology, 68(1), 1-18. https://doi.org/10.1093/sysbio/syy035 Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). The publication may also be distributed here under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license. More information can be found on the University of Groningen website: https://www.rug.nl/library/open-access/self-archiving-pure/taverne- amendment. Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 30-09-2021

University of Groningen Inferring Diversification Rate

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: University of Groningen Inferring Diversification Rate

University of Groningen

Inferring Diversification Rate Variation From Phylogenies With FossilsMitchell, Jonathan S; Etienne, Rampal S; Rabosky, Daniel L

Published in:Systematic biology

DOI:10.1093/sysbio/syy035

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

Document VersionPublisher's PDF, also known as Version of record

Publication date:2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):Mitchell, J. S., Etienne, R. S., & Rabosky, D. L. (2019). Inferring Diversification Rate Variation FromPhylogenies With Fossils. Systematic biology, 68(1), 1-18. https://doi.org/10.1093/sysbio/syy035

CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

The publication may also be distributed here under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license.More information can be found on the University of Groningen website: https://www.rug.nl/library/open-access/self-archiving-pure/taverne-amendment.

Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

Download date: 30-09-2021

Page 2: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 1 1–18

Syst. Biol. 68(1):1–18, 2019© The Author(s) 2018. Published by Oxford University Press, on behalf of the Society of Systematic Biologists. All rights reserved.For permissions, please email: [email protected]:10.1093/sysbio/syy035Advance Access publication May 18, 2018

Inferring Diversification Rate Variation From Phylogenies With Fossils

JONATHAN S. MITCHELL1,2, RAMPAL S. ETIENNE3, AND DANIEL L. RABOSKY1,∗1Museum of Zoology and Department of Ecology and Evolutionary Biology, University of Michigan, 1105 North University Avenue,

Ann Arbor, Michigan 48109 USA; 2Department of Biology, West Virginia University Institute of Technology, 410 Neville Street, Beckley, WV 25801, USA;and 3Groningen Institute for Evolutionary Life Sciences, University of Groningen, Box 11103, 9700 CC Groningen, The Netherlands∗Correspondence to be sent to: Museum of Zoology and Department of Ecology and Evolutionary Biology, University of Michigan,

1105 North University Avenue, Ann Arbor, MI 48109, USA;E-mail: [email protected].

Received 7 March 2017; reviews returned 8 May 2018; accepted 9 May 2018Associate Editor: Richard Ree

Abstract.—Time-calibrated phylogenies of living species have been widely used to study the tempo and mode of speciesdiversification. However, it is increasingly clear that inferences about species diversification—extinction rates in particular—can be unreliable in the absence of paleontological data. We introduce a general framework based on the fossilized birth–death process for studying speciation–extinction dynamics on phylogenies of extant and extinct species. The model assumesthat phylogenies can be modeled as a mixture of distinct evolutionary rate regimes and that a hierarchical Poisson processgoverns the number of such rate regimes across a tree. We implemented the model in BAMM, a computational frameworkthat uses reversible jump Markov chain Monte Carlo to simulate a posterior distribution of macroevolutionary rate regimesconditional on the branching times and topology of a phylogeny. The implementation, we describe can be applied topaleontological phylogenies, neontological phylogenies, and to phylogenies that include both extant and extinct taxa. Weevaluate performance of the model on data sets simulated under a range of diversification scenarios. We find that speciationrates are reliably inferred in the absence of paleontological data. However, the inclusion of fossil observations substantiallyincreases the accuracy of extinction rate estimates. We demonstrate that inferences are relatively robust to at least someviolations of model assumptions, including heterogeneity in preservation rates and misspecification of the number ofoccurrences in paleontological data sets. [BAMM; comparative methods; extinction; fossils; macroevolution; speciation.]

Variation in rates of speciation and extinction con-tribute to striking disparities in species richness amongdifferent clades of organisms. The study of macroe-volutionary rates was pioneered in the fossil record,with the application of phenomenological birth–deathmodels to patterns of diversity through time (Raupet al. 1973; Sepkoski 1978; Raup 1985). Phylogenies ofextant species are widely used to estimate rates of bothspeciation and extinction (Nee et al. 1994; Nee 2006;Alfaro et al. 2009; Morlon 2014). However, the integrationof paleontological and neontological data will likelyprovide more accurate insights into macroevolutionarydynamics than molecular phylogenies alone (Quentaland Marshall 2010; Liow et al. 2010a; Stadler 2013;Rabosky 2016). Indeed, a recent simulation study hasargued that macroevolutionary rates inferred using onlya tree topology and fossil occurrence dates are morereliable than rates inferred with an extant-only tree andbranch lengths (Didier et al. 2017).

Inferences on macroevolutionary patterns frequentlydiffer between paleontological and molecular phylogen-etic data (Hunt and Slater 2016). Fossil data routinelysupport high levels of extinction and large fluctuationsin standing richness (e.g., Alroy 1999, 2008, 2010; Liowet al. 2010a; Quental and Marshall 2010; Sheets etal. 2016). However, extant-only analyses rarely findsupport for high extinction rates (Nee 2006; Purvis2008, but see Etienne et al. 2012), and numerous studieshave addressed potential reasons for this discrepancy(Rabosky 2009, 2010, 2016; Morlon et al. 2011; Etienne andRosindell 2012; Etienne et al. 2012; Rosenblum et al. 2012;Beaulieu and O’Meara 2015). Directly incorporating

fossil data into phylogenetic rate estimation shouldfacilitate more robust tests of the role of extinction inmacroevolutionary dynamics (Sakamoto et al. 2016). Anexplicitly phylogenetic approach to macroevolutionaryrates in the fossil record will provide insights into ratevariation among lineages and will also facilitate thestudy of diversification dynamics in clades that aretoo poorly preserved for subsampling-based approaches(e.g., Alroy 1999).

In this article, we describe a general framework formodeling complex speciation–extinction dynamics onphylogenetic trees that include both paleontological andneontological data. The model is based on the fossilizedbirth–death process (Stadler 2010; Didier et al. 2012,2017; Heath et al. 2014) and includes parameters forspeciation, extinction, and fossilization rates. The birth–death process for reconstructed phylogenetic trees (Neeet al. 1994) is a special case of the fossilized birth–deathprocess where the preservation rate is equal to zero(Stadler 2010). As such, this model can be applied topaleontological phylogenies (extinct taxa only), neonto-logical phylogenies of living species (extant taxa only),and phylogenies that include any combination of extantand extinct species.

The novel feature of our implementation, relative toprevious work on diversification rates from phylogeniesand fossils (Stadler 2010; Didier et al. 2012; Heath etal. 2014), is the ability to capture complex patternsof among-lineage variation in speciation and extinc-tion rates. The model assumes that phylogenies areshaped by a countably finite set of distinct evolutionaryrate regimes, and that the number of such regimes

1

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 3: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 2 1–18

2 SYSTEMATIC BIOLOGY VOL. 68

is governed by a hierarchical Poisson process. Thismodel is an extension of the multiprocess diversificationmodel currently implemented in the BAMM softwarepackage for macroevolutionary rate analysis (Rabosky2014; Rabosky et al. 2017). We note that Sakamoto et al.(2016) developed a flexible regression-based frameworkfor estimating heterogeneity in diversification rates fromfossil phylogenies, but their approach does not rely ona formal process-based model of diversification andpreservation.

We first describe the analytical formulation of themodel and state its assumptions. We then describe ourimplementation within the BAMM software frameworkand discuss its relationship to previous BAMM versions(Rabosky et al. 2017). To evaluate the performance ofthe model, we simulated phylogenies under a range ofdiversification scenarios, and then subjected each sim-ulated phylogeny to stochastic fossilization to generatea tree similar to what researchers would typically use(an “observed” phylogenetic tree with an incompletelyfossilized history). We analyzed each phylogeny withBAMM and compared the BAMM-inferred rates andshift configurations to the known (simulated) valuesto assess the accuracy of parameter estimates. We findthat the inclusion of even sparse paleontological datain phylogenies dominated by extant lineages improvesthe accuracy of both speciation and extinction rateestimates. This improvement is especially pronouncedin estimation of extinction rates. Although our modelmakes several simplifying assumptions about the natureof fossilization, we demonstrate that rate inferencesare reasonably robust to temporally-heterogeneity inpreservation rates.

METHODS

DataThe data consist of a time-calibrated phylogenetic tree

(which may include both extinct and extant lineages) anda count of the number of stratigraphically-distinct fossiloccurrences associated with the tree. Each extinct lineagein the phylogeny should be represented by a branch thatterminates at the time of the last occurrence for the taxon.Note that true extinction events are not observed directly,and the last occurrence time is expected to precede thetrue extinction time of a taxon. The number of fossiloccurrences is best thought of as an estimate of thenumber of fossils that could be, in principle, assignedto lineages present in the tree. Fossil occurrences ofindeterminate nature and that are potentially associatedwith lineages not present in the tree (e.g., “Theropoda,indeterminate”) are not accommodated in this modelingframework and are excluded from the count of occur-rences; the likelihood calculations only apply to fossilsthat can be assigned explicitly to branches in the tree. Putanother way, fossil occurrences should only be countedif the species-level taxon to which they are assignedis included in the tree. However, because the currentimplementation assumes homogeneity of preservation

through time and among lineages, the calculations donot actually use stratigraphic and phylogenetic contextfor individual fossil observations. As explained below,this assumption of time-homogeneous preservationmeans that the likelihood of the phylogeny with fossildata is invariant with respect to the location of the fossilson the tree.

Likelihood of a Phylogeny with Fossils and Rate-ShiftsThe fundamental operation in the analysis of diver-

sification rate shifts on phylogenetic trees involvescomputing the likelihood of the phylogeny under a fixedconfiguration of lineage types. Each “type” of lineageis a characterized by a distinct and potentially time-varying rate of speciation (�) and extinction (�). Twolineages i and j are said to be of the same type if �i(t)=�j(t) and �i(t)=�j(t) everywhere in time. Thus, for agiven point in time t, lineages of the same type willhave precisely the same distribution of progeny lineagesat some time t+�t in the future. For the constant-ratebirth–death process, all lineages are assumed to be of thesame type. For the Binary State Speciation and Extinction(BiSSE) process (Maddison et al. 2007), phylogenies areassumed to comprise a mixture of two types of lineages,and the character states at the tips of the tree can beviewed as labels for lineage types. The transition pointsbetween types are unknown in the BiSSE process, and thelikelihood of a given set of labeled types can be computedby integrating over all possible transitions that couldhave given rise to the observed data.

In contrast to BiSSE, the calculations implemented inBAMM and other rate-shift models (Alfaro et al. 2009;Morlon et al. 2011; Etienne and Haegeman 2012) assumethat we have mapped lineage types across all branchesof the phylogeny and not merely to the tips. We define a“mapping” of rate parameters as an association betweeneach point on a phylogeny (e.g., segment of a branch) anda particular lineage type. Additional steps may be usedto optimize the number, location, and parameters of themapped types, including maximum likelihood (Alfaroet al. 2009; Morlon et al. 2011; Etienne and Haegeman2012).

Likelihood calculations for the fossilized birth–deathprocess were described by Stadler (2010), and our generalapproach extends these calculations to a phylogeny withmultiple types of lineages. The fossilized process differsfrom the reconstructed evolutionary process (Nee etal. 1994) because we must also account for the lineagefossilization rate �. The new calculations describedbelow are implemented in BAMM v2.6.

The derivation of the likelihood for rate-shift modelsfollows the logic described in Maddison et al.’s (2007)description of the BiSSE model. In a given infinitesimalinterval of time �t, the state-space for the processincludes five events that can potentially change theprobability of the data: a lineage may undergo arate shift, become preserved as fossil, become extinct,speciate, or it may undergo none of these events. Tocompute the likelihood of the tree and a set of parameters

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 4: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 3 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 3

(e.g., a mapping of rate shifts), we assume that noother rate shifts have occurred, thus dropping one ofthe possible events (the occurrence of an unmappedrate shift) from the state space for the process. Assuch, the likelihood is conditioned on the set of shiftsthat have been placed on the tree. This assumptionof rate-shift models has been criticized in the recentliterature (Moore et al. 2016). However, it is unclearhow we might accommodate unobserved rate shifts,even in principle, as there is very little informationin the data or elsewhere that can yield informationabout the probability distribution of unobserved rateparameters. Most importantly, we demonstrate thatthe simplifying assumptions that underlie the BAMMcalculations yield robust inference on evolutionary rates,even when simulated phylogenies used for validationcontain many such unobserved rate shifts.

We assume that preservation events are mutuallyexclusive with respect to speciation and extinctionevents. Hence, a lineage cannot leave a recoverablefossil observation during the precise moment of lineagesplitting or lineage extinction. This interpretation ofthe preservation rate ensures that the model we haveimplemented is mathematically identical to the modeldescribed by Stadler (2010), for the special case wherespeciation and extinction rates do not vary amonglineages. This interpretation of preservation is alsoconsistent with paleontological practice: fossil observa-tions are always assigned to single lineages. Even inthe relatively high-resolution marine microfossil record,paleontologists do not assume that the data capture theprecise instant of lineage splitting (Benton and Pearson2001). In this article, we assume that lineage preservationis a time-homogeneous process, but it is straightforwardto relax this assumption.

We measure time t backwards from some arbitrarystarting time, t0. The start time, t0, need not correspondto the present. Let D(t) be the probability density that alineage that is extant some time t before t0 gives rise tothe observed data at the present time, and let E(t) be thecorresponding probability that a lineage that is extantat time t goes extinct, along with its descendants, beforetime t0. Following Stadler (2010), the master equationsgoverning the likelihood of the data are

dDdt

=−(�+�+�)

D(t)+2�D

(t)E

(t)

(1)

anddEdt

=�−(�+�+�)

E(t)+�E

(t)2 (2)

E(t) describes the probability that an independentlineage originating at time t goes extinct, along with allof its descendants, before some future time t0. One canalso view this probability as the probability of “non-observation” of a lineage and all of its descendants;a lineage and its descendants might not be observedbecause they have gone extinct, and/or because we havefailed to sample them. To compute the likelihood of thephylogeny with mapped rate shifts, we require solutions

to these equations for arbitrary t and initial values t0, D0,and E0. For E(t), the solution is (Stadler 2010)

E(t)=

�+�+�+c1e−c1�t(1−c2

)−(1+c2

)e−c1�t(1−c2

)+(1+c2

)

2�(3)

where

c1 =∣∣∣∣√(�−�−�)2 +4��

∣∣∣∣ (4)

and

c2 =−�−�−�−2�(1−E0

)

c1. (5)

For D(t, t0), we have

D(t)= 4D0

2(

1−c22

)+e−c1�t

(1−c2

)2 +ec1�t(1+c2

)2(6)

which follows immediately from Stadler (2010) The-orem 3.1, under the substitution D0 =�. In theSupplementary Material, we provide a derivation of D(t)for arbitrary D0.

Calculations are performed by solving D(t) and E(t)recursively on the phylogeny from the tips to the root,combining D(t) probabilities at internal nodes (discussedbelow) and continuing down the tree in a rootwardsdirection. For lineages that are extant, we have initialconditions at time t0 = 0 of D0 =� and E0 = 1−�,where � is the probability that an extant lineage hasbeen sampled (e.g., the sampling fraction; FitzJohn etal. 2009). In the absence of perfect preservation, thelast fossil occurrence of a given taxon is not the trueextinction time. For a lineage with a last occurrence attime ti, we generally know that the lineage and all ofits descendants went extinct sometime between ti andt0 = 0. Hence, for terminal lineages with non-extant tips,we initialize both E0 and D0 with E(ti|t0 = 0), or theprobability that a single lineage that is extant at timeti will have zero observed descendants prior to andincluding the present day (t= 0). This initialization forD0 accounts for the probability that a non-extant tipgoes extinct, or is unsampled, on the time interval (ti,0). At internal nodes, we combine the densities from theright DR(t) and left DL(t) descendant lineages, such thatD(t) immediately rootwards of the node is computedas D(t)=�(t)DR(t)DL(t). The full likelihood must alsoaccount for all Z observed fossil occurrences, which isaccomplished by multiplying the probability density ofthe tree by �Z.

There is one difference between the implementationof the model described here (implemented in BAMMv2.6) and the default calculation scheme implementedin the most recent previous version, BAMM v2.5. Asdiscussed in Rabosky et al. (2017), there are severalalgorithms for computing the likelihood of a specifiedmapping of rate regimes on a phylogenetic tree. Thesealgorithms, which we have termed “recompute” and“pass-up” (Rabosky et al. 2017), differ in how theanalytical E(t) equation is initialized for calculations on

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 5: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 4 1–18

4 SYSTEMATIC BIOLOGY VOL. 68

individual branch segments. The precise sequence of cal-culations associated with these approaches is describedin the Supporting Information (Section S2.4 available onDryad at https://doi.org/10.5061/dryad.50m70) fromRabosky et al. (2017). Herein, we expand BAMM toinclude the “recompute” algorithm for computing thelikelihood, which was not implemented as the defaultcalculation scheme in BAMM v2.5. The process underconsideration in the present article can yield phylogeniesthat have gone extinct in their entirety (e.g., fossil taxaonly), and we can thus avoid conditioning on survival ofthe process as well as unresolved theoretical problemsassociated with this conditioning. We thus perform allcalculations using both the “recompute” and “pass-up” algorithms; we present “recompute” results in themain text, but the corresponding results using “pass-up” were virtually identical and are presented in theSupplementary Information available on Dryad.

As noted above, our use of these equations formodeling diversification dynamics implicitly conditionsthe likelihood on a finite and observed set of lineagetypes: the likelihood does not account for new typesof lineages that may have evolved on lineages with nofossilized or extant descendants (Moore et al. 2016).This assumption applies to all methods that purportto compute the likelihood of a phylogeny under aspecified mapping of rate shifts (Alfaro et al. 2009,Morlon et al. 2011, FitzJohn 2010, Etienne and Haegeman2012). Indeed, this assumption applies indirectly toall diversification methods: if there are no (or veryfew) unobserved shifts, the likelihood computed byBAMM is approximately correct. If “reality” containsmany such unobserved shifts, then the likelihood (andcorresponding rate estimates) from BAMM will bebiased, because rate shifts will essentially have servedas an unaccommodated source of lineage extinction. Butif unobserved rate shifts are sufficiently common (inreal data) as to bias BAMM, then they are problematicfor all other current methods and models as well.As pointed out by Rabosky et al. (2017), the problemof unobserved rate shifts does not disappear simplybecause one assumes a theoretically complete model thatignores their existence or by sampling their parametersfrom a prior (assumed) distribution.

Implementation in BAMMThe model described above is implemented as an

extension to the BAMM software (v2.6). BAMM v2.6expands the basic diversification model implemented inearlier versions of BAMM by incorporating parametersthat specify the total observation time of the tree(“observationTime”, Tobs: time from root when the treeis observed; this parameter is equal to the root agefor trees with extant lineages), the total number ofobserved fossil occurrences within the clade analyzed(“numberOccurrences”), and parameters that controlthe prior distribution of, and update frequency for, thepreservation rate� (“updatePreservationRate”, “preser-vationRatePrior”, and “preservationRateInit”).

Tobs can have a large impact on the overall modelfit, and warrants additional explanation. The modelassumes that non-extant tip lineages can go extinctanywhere on the time interval that extends from theirlast occurrence to Tobs. Consider the scenario wherewe are modeling diversification dynamics in an extinctclade of dinosaurs that originated 200 million yearsbefore the present (Ma) and where the last occurrenceof that clade is 90 Ma. If we set Tobs to equal the presentday (0 Ma), the model will allow the possibility thatthe last dinosaur from the clade went extinct sometimebetween 90 and 0 Ma. In practice, we suggest choosingthe observationTime as the earliest possible time that alllast-occurrence taxa are believed to have gone extinct.In the dinosaur example, a reasonable setting for Tobswould be the KPg boundary at 66 Ma, by which time allnon-avian dinosaur lineages had become extinct.

The number of occurrences (parameter “numberOc-currences” in BAMM) represents the total number offossil occurrences that can be assigned to observedlineages in the clade under consideration and is notsimply the number of extinct tips in the phylogeny. Itis necessarily the case that a phylogeny with k extincttips has at least koccurrences: one for the last occurrenceof each extinct lineage; however, there can be multipleoccurrences for each extinct and extant lineage. Thereare several possible interpretations of a fossil occurrencein the BAMM framework; we address this issue atlength in the Discussion section. Occurrences associatedwith fossil taxa that are not present in the tree shouldnot be used, nor should taxonomically indeterminatematerial (e.g., fragmentary fossil material that cannotbe taxonomically identified below the genus level).In general, we consider an “occurrence” to repres-ent stratigraphically-unique species-level observations.Hence, a locality with 1000 individuals of a single taxonrepresenting a single depositional event (e.g., a massburial event) would constitute a single occurrence. Inpractice, we suggest analyzing a data set with severaldifferent values for the occurrence count to ensure thatresults are robust (see Discussion section).

Description of Tree SimulationWe developed software to simulate phylogenies

under a Poisson process that includes shiftsto new macroevolutionary rate regimes(https://github.com/macroevolution/simtree). Eachsimulation was initialized with two lineages at time t= 0with � and� drawn from exponential distributions withmeans of 0.2 and 0.18, respectively. We assumed thatshifts to new rate regimes occurred with rate �= 0.02.Waiting times between events (speciation, extinction,rate shift) were drawn from an exponential distributionwith a rate equal to the sum of the rate parameters(�+�+�). For rate shift events, new values of � and �were drawn from the same exponential distributions asthe root regime. This procedure simulates a phylogenywith stochastic rate shifts; within a given rate regime,� and � were constrained to be constant in time.

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 6: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 5 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 5

Simulations were run for 100 time units, and treeswith more than 5000 or fewer than 50 lineages at t=100 were discarded, resulting in a total of 200 trees foranalysis. In addition, each tree was constrained to haveat least one rate shift in the true, generating history. Thissimulation procedure can generate phylogenies withunobserved rate shifts when shifts occur in unsampledextinct clades (see below). Conversely, the true rootregime may be unobserved and all observed tips mayfall within one of the derived regimes. As discussed inthe section below on fossilization, any rate shift thatleads to an extinct clade with no fossil observationswill be unobserved, thus enabling us to assess whetherMoore et al.’s (2016) concerns about unobserved rateshifts have consequences for inference in practice.

Description of Fossilization ProcedureWe created pseudo-fossilized trees by stochastically

placing fossils along the branches of each simulated tree(Fig. 1) using per-lineage preservation rates (�) of 0, 0.01,0.1, or 1.0. Speciation rates in our simulations rangedfrom 0.003 to 1.7 lineages per unit time, with 99% ofthe regimes having a speciation rate above 0.015; 80%exceeded 0.1. As such, speciation events were generallymore frequent than fossilization events, except underthe high preservation (�= 1.0) simulation scenario. Allextant tips, and tips with at least one fossil occurrence,were retained while tips (and clades) lacking fossiloccurrences were dropped. In empirical data sets, thetrue extinction time of an extinct lineage is unknown,and the time of the last (most recent) fossil occurrenceis used. For comparability with empirical data from thefossil record, extinct tips with fossils were truncated tothe time of the most recent fossil occurrence (e.g., thebranch ends at the last occurrence, and any subsequenthistory for the lineage is necessarily unobserved). Werefer to these pseudo-fossilized trees as “degraded,”because the stochastic fossilization history necessarilydiscards information about the true lineage history.

As pointed out by Moore et al. (2016), the modelimplemented in BAMM assumes that rate shifts do notoccur on unobserved branches. However, our simulationand fossilization procedure allows rate shifts to occur onextinct clades with no fossil observations. Hence, our fos-silization procedure frequently leads to unobserved rateregimes and enables us to test the practical significanceof this core BAMM assumption.

BAMM AnalysesWe approximated the posterior distribution of rate

shift configurations for each degraded tree by runninga Markov chain Monte Carlo (MCMC) simulation inBAMM v2.6 for 50 million generations, with the first5 million discarded as burnin. Each tree was analyzedassuming a single, time-constant rate of preservation.The model prior (mean of the geometric prior distribu-tion on the number of shifts) was set to 200 for all runs

a

b

FIGURE 1. Diagram showing a complete evolutionary treewith unobserved (gray) and observed (black) evolutionary history.Fossils are indicated by open circles; labeled circles (a, b) denote rateshifts on individual branches. Tree contains three extant tips, threeobserved extinct tips, and eight unobserved extinct tips. Extinct taxa arenecessarily placed at the time of their last fossil occurrence. Rate shift(a) is unobserved in phylogenies that include fossil data, as there areno extant descendants or fossil occurrences on the subtree descendedfrom the shift. The simulation protocol used in this study involvedgenerating phylogenies for the complete process, then simulating fossiloccurrences on the tree and removing unobserved (gray) evolutionaryhistory.

(after Mitchell and Rabosky 2016), while the ratesof the exponential priors on � and � were scaledusing the setBAMMpriors() function in BAMMtools(see Supplementary Material available on Dryad fordetails). Mitchell and Rabosky (2016) found that infer-ences on the number of shifts were generally robustto the choice of model prior, but that use of a liberalmodel prior facilitated convergence of the MCMCsimulation. We performed each analysis using both“recompute” and “pass-up” algorithms for computingthe likelihood, as described by Rabosky et al. (2017).All computer code, data files, and result files areavailable through the Dryad digital data repository(https://doi.org/10.5061/dryad.50m70).

Performance AssessmentTo assess the performance of our model and BAMM

implementation, we quantified four fundamental attrib-utes of interest: 1) the estimated preservation rate, 2)the speciation and extinction estimates inferred for each(true) rate regime, 3) the inferred number of macroevolu-tionary regimes on the tree, and 4) the location of inferredrate shifts. We compared the posterior distribution ofthe estimated preservation rate � for each degraded

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 7: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 6 1–18

6 SYSTEMATIC BIOLOGY VOL. 68

tree to the generating value. For diversification rateswe found the mean inferred speciation and extinctionrates for every true regime with at least two tips (bothextinct and extant), and compared these mean valuesto the true values of speciation (�), extinction (�) andrelative extinction (�/�). We predicted that BAMMwould estimate rates more accurately for regimes withlarger numbers of taxa; we thus assessed BAMM’saccuracy with explicit reference to the number of tips ineach regime. This allowed us to assess how well BAMMestimates rates for regimes that differ in size.

We determined the best-supported number of shiftsfor each tree by using stepwise model selection withBayes factors (Mitchell and Rabosky 2016). Using the pos-terior distribution of shifts for each tree, we comparedthe lowest complexity model (e.g., smallest numberof shifts) to the next-most complex model. If themore complex model was supported by Bayes factorevidence of at least 20 relative to the simpler one, itwas provisionally accepted and the process repeatedwith models of increasing complexity until the Bayesfactor threshold of 20 was no longer exceeded. Wethen compared the best-supported number of shifts onthe tree to the true number of preserved shifts (seeFig. 1).

Finally, we looked at how accurately BAMM inferredthe locations of shifts along the tree. Location accuracyis difficult to quantify, as many simple metrics haveundesirable properties. For example, the distance alongthe branches between a true and inferred shift is difficultto interpret unless the number of inferred shifts is exactlyequal to the number of true shifts. Branch lengths furtherconfound location accuracy as shifts along shorterbranches are expected to be more “smeared” acrossadjacent branches relative to shifts on longer branches(see extensive discussion of this topic in the supplementto Mitchell and Rabosky 2016). We thus assessed locationaccuracy using several complementary metrics.

The first metric we used for location accuracywas the marginal posterior probability of at leastone rate shift on all branches that included a (true)rate shift. We assessed these marginal probabilitieswith respect to both the magnitude of the shift(e.g., size of the rate increase) as well as the num-ber of tips descending from that shift, with theexpectation that shifts with many descendent tipsand/or with a large change in rates would havehigher marginal posterior probabilities. As pointedout by Rabosky (2014), focusing on these marginalprobabilities can be misleading if the evidence fora rate shift is “smeared” across a set of adjacentbranches.

As an alternative metric of location accuracy, we usedthe method described by Mitchell and Rabosky (2016)to assess how well BAMM partitions tips into theircorrect rate classes. This approach can be motivated bynoting that any placement of k rate shifts on a phylogenypartitions a tree into at most k+ 1 subsets of tips withdistinct rate regimes (k+ 1, not simply k, due to thepresence of a root regime that is distinct from the rate

shifts). For a phylogeny with N taxa and a known (true)mapping of diversification regimes, there is an N x Nmatrix where each element Ci,k describes whether thei’th and k’th taxa are assigned to the same or differentrate regimes (the set of tips partitioned into a particularrate regime was termed a “macroevolutionary cohort” inRabosky 2014). Elements of the matrix Ci,k take a valueof 1 (if taxa i and k are in the same rate regime), or 0 (ifi and k are in different rate regimes). Following BAMManalysis, we obtain a matrix D where Di,k represents theposterior probability that taxa i and k are assigned to thesame rate regime. To assess shift location accuracy, wecomputed one minus the average of the absolute valueof the difference between the true pairwise partitioningmatrix C and the BAMM-estimated matrix D, or

1− 2N

(N−1

)N∑

k=2

k−1∑i=1

∣∣Ci,k −Di,k∣∣ (7)

An overall value of 1.0 can only be obtained if allpairs of taxa are correctly partitioned in the correctconfiguration of rate regimes in 100% of samples fromthe posterior. This metric of location accuracy willpenalize shift configurations that include too manyshifts (by separating taxa that should be in the sameregime) and that include too few shifts (by unitingtaxa that should be in separate regimes). However,the topology of the tree may make certain partitionsmore likely by chance alone. To quantify this baselineaccuracy, we computed the partitioning accuracy thatwould be expected by chance if shifts were distributedrandomly throughout the tree under a uniform priorwith respect to the total tree length. Under a uniformprior, the probability of a shift on a given branchbi is simply bi/T, where bi is the length of the i’thbranch and T is the tree length. We conditioned therandomization on the simulated shift count, such thateach randomized sample contained exactly the samenumber of shifts as the corresponding sample fromthe posterior. After randomizing the placements ofshifts from all samples in the posterior, we computedthe accuracy of random partitions for each sampledgeneration using the same pairwise distance matrixapproach described above.

Sensitivity of the Method to Model ViolationsWe tested sensitivity of the framework described

here to four qualitatively distinct violations of theassumptions of the underlying inference model. Theseviolations included: 1) violation of time-homogeneouspreservation potential; 2) misspecification of the truenumber of fossil occurrences in the data; 3) presence of amass extinction in the data; and 4) diversity-dependentdiversification. In real data sets, preservation rates canvary substantially even over relatively short timescales(Foote and Raup 1996; Holland 2016). We violated theassumption of time-constant preservation by applyinga “stage-specific” preservation pattern to the simulatedphylogenies. We use “stage” in this context solely to

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 8: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 7 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 7

denote an arbitrary temporal span with a potentiallydistinct preservation rate. We divided each simulatedtree into ten equally-sized temporal bins (each ten timeunits long, mirroring the 10 myr bins used in somelarge-scale paleodiversity studies), where each bin wasassigned an independent preservation rate drawn froma uniform (0.01, 0.9) distribution. We then simulatedfossilization histories using these preservation ratesand pruned unobserved lineage histories as describedabove to generate a set of degraded trees. We analyzedthese trees using BAMM v2.6, which assumes time-homogeneous preservation, and we tested the accuracyof inference using the same metrics described above.Given that the simulated trees involve rate shifts, thissimulation approach allows clades to radiate in periodsof higher or lower preservation, and to persist intosubsequent time bins with distinct and decoupledpreservation parameters.

Our implementation assumes that users can specify ameaningful number of fossil occurrences for the clade.Fossil occurrences are not placed at specific points on thetree, but rather inform the model about the total numberof fossil occurrences associated with the focal clade. Thecurrent implementation thus assumes fossil occurrences(with the exception of last occurrences) are distributedacross the tree under a homogeneous Poisson process,and estimates a tree-wide rate accordingly. However, forreal data sets, the “true” number of occurrences is likelyimpossible to know with precision (even in principle;see Discussion section). To assess how misspecificationof the number of fossil occurrences may bias inference,we simulated 100 constant rate phylogenies, with higherrates than the variable-rate tree simulations (mean�= 0.9, mean �= 0.45) and imposed a fossilizationhistory with � of 0.1. Given the stochastic nature ofthe fossilization simulation, the fossilized trees variedin the number of extinct lineages observed. Note thatbecause an extinct tip is only observed if at least onefossil occurs on that branch, each fossilized tree with(observed) extinct lineages has a minimum number offossil occurrences equal to the number of (observed)extinct tips (FMIN). We then calculated the differencebetween this per-tree minimum and the true number(FTRUE − FMIN = FGAP), and analyzed the tree usingBAMM v2.6 with the observed number of occurrencesset to: 1) FMIN, 2) FMIN+ 0.25 * FGAP, 3) FMIN+0.5 * FGAP, 4) FTRUE, 5) FTRUE+ FGAP, and 6) 10 *FTRUE. Overestimation of the true number of fossilsmay seem unlikely, but this could arise in practice dueto taxonomic error (e.g., incorrect assignment of fossilobservations to a specific taxon) or to the inclusion ofindeterminate fossil material in the analysis. Scaling thedegree of misspecification to the true and minimumnumber of fossils created roughly comparable degreesof misspecification across trees with large differences inthe number of observed extinct lineages.

We estimated the bias for each of these misspe-cified fossil occurrence values by dividing the inferredspeciation and extinction rates across the entire tree(as inferred using the getCladeRates function from

BAMMtools) by the true values used to simulate the tree.Our discussion below pertains to bias in rates resultingfrom misspecification of the number of occurrences,not bias due a particular magnitude of preservation(e.g., “low” or “high” preservation). We predictedthat overestimates of fossil occurrences should leadto underestimates of evolutionary rates. Essentially,as the number of fossil occurrences is increased, theinferred preservation rate will be higher; this increasedpreservation rate, in turn, will decrease the amount ofunobserved evolutionary history that is assumed to haveoccurred. By reducing the inferred unobserved historythrough artificial inflation of the preservation rate, weunderestimate the number of events that have occurred(speciation or extinction), and the corresponding rateestimates are biased downwards. When the observednumber of fossil occurrences is lower than the truenumber, however, we will tend to overestimate both theunobserved evolutionary history and the correspondingdiversification rates.

We also tested the impact of diversification historiesthat violate the assumptions of the BAMM frameworkusing two scenarios. First, we asked how the recentextinction of a large clade may bias the estimates fromBAMM. To simulate this scenario, we chose a single,large, extant clade on each simulated phylogeny andpruned each extant branch back to the time of themost recent fossil occurrence (or dropped it entirely,if it was never observed in the fossil record). For the2nd scenario, we simulated trees under a diversity-dependent process, and inferred rates from that tree bothwith and without fossil data. Further details on thesescenarios are presented in the Supplementary Materials(Section S8 available on Dryad).

Empirical AnalysisFor comparison purposes, we include an empirical

analysis using BAMM of a recently published 138-tipphylogeny of horses (Cantalapiedra et al. 2017). Wetime-scaled the horse tree using the same approachas Cantalapiedra et al. (2017), which tends to homo-genize branches by assuming all branch lengths arepulled from a single distribution defined by a spe-ciation, extinction and preservation rate (see Bapst2013; and the Supplementary Materials available onDryad). We compared the BAMM results to thoseobtained using PyRate (Silvestro et al. 2014a,b), aBayesian inference framework that allows estimationof diversification parameters and rate shifts from theoccurrence data alone. We downloaded 1495 fossiloccurrences (see Discussion) for the 138 species in thephylogeny from the Paleobiology Database (Alroy etal. 2001) using the R interface “paleobioDB” (Varelaet al. 2015). We simulated the posterior distribu-tion of rate shift configurations using BAMM with20,000,000 generations of MCMC sampling. The PyR-ate analysis used 1,000,000 generations of MCMCsampling.

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 9: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 8 1–18

8 SYSTEMATIC BIOLOGY VOL. 68

RESULTS

Simulated trees varied from 50 to 4578 tips (mean= 770), and degrading the trees reduced the observednumber of lineages to an average size of 589, 616, and698 tips for the lowest to highest preservation rates.Because multiple fossilization histories were imposedon each fixed evolutionary history, the number of fossiloccurrences scales directly with the preservation rate(mean numbers of fossils of 25.4, 255.6, and 2549.9 forthe lowest to highest preservation rates). After 50 milliongenerations, the MCMC chains for at least 95% of datasets in each fossilization scheme had reached our targetconvergence criterion, with effective sample sizes for thenumber of shifts at over 200. MCMC simulations thatfailed to meet this convergence criterion are neverthelessincluded in all following analyses to avoid biasing thesubsample by including only trees that showed goodMCMC performance. Results based on all trees, eventhose that failed to achieve the desired effective samplesizes, are nearly identical to the results presented belowand are shown in the Supplementary Material availableon Dryad.

For each rate regime in each simulated tree, wecompared the estimated rates with the true rates inthe generating process. Estimated speciation rates werereasonably well correlated with true rates across allpreservation levels. Across all rate regimes and pre-servation scenarios the correlation between true andestimated speciation rates was 0.49, although 70.9%of these regimes contained fewer than five tips. It isunlikely that BAMM would have had sufficient powerto infer such small regimes (Rabosky et al. 2017), andestimated rates for most will tend to equal the rateof the parent process. For regimes with five or moretips, the correlation increases to 0.77 and 0.86 for theextant-only and high-preservation (�= 1.0) scenarios,with all other preservation scenarios falling betweenthese values (Fig. 2, left column). For regimes with 20or more tips, these correlations rise to 0.90 and 0.97,respectively (Fig. 3, left column). The regression slopesare close to the expected value of one, but do notshow a consistent increase with preservation (extant-only: slope5+ = 0.91 ± 0.01 s.e. and slope20+ 1.01 ± 0.01s.e. for regimes with 5+ tips and regimes with 20+tips, respectively; �= 0.01: slope5+ 0.90 ± 0.01 s.e. andslope20+ 0.99 ± 0.01 s.e.; �= 0.1: slope5+ 0.86 ± 0.01s.e. and slope20+ 0.94 ± 0.01 s.e.; �= 1: slope5+ = 0.88± 0.01 s.e. and slope20+ = 0.96 ± 0.01 s.e.). Speciationrate correlations improve as regimes with an increasingminimum number of tips are considered, regardless ofpreservation rate (Fig. 4).

Estimates of the extinction rate (�) are substantiallyinfluenced by the preservation rate. Across all regimesand preservation rates, the mean correlation betweentrue and estimated regime-specific rates was 0.14. Asfor speciation, however, these correlations improvesubstantially when we exclude the 70.9% of rate regimesthat included fewer than five tips. For regimes withfive or more tips, correlations ranged from 0.30 for

extant-only trees to 0.72 for the �= 1.0 scenario (Fig. 2).With 20 or more tips, these correlations increase to0.49 and 0.91, respectively (Fig. 3). As with speciation,the slopes do not vary consistently with preservation,although the error decreases with increasing preserva-tion (extant-only: slopeall = 0.54 ± 0.02 and slope20+ =1.20 ± 0.05 for all and regimes with 20+ tips; �=0.01: slopeall = 0.47 ± 0.02 and slope20+ = 0.95 ± 0.05;�= 0.1: slopeall = 0.38 ± 0.01 and slope20+ = 0.78± 0.02; �= 1: slopeall = 0.50 ± 0.01 and slope20+ =0.95 ± 0.01).

The relative extinction rates (�/�) also improved withincreasing preservation, with the estimates from extant-only simulations only weakly correlated with the truerelative extinction level (r= 0.41 for 5+ tips, 0.53 for 20+tips) and again increasing with increased preservation(r= 0.67 for regimes with 5+ tips; 0.87 for regimeswith >20 tips; Figs. 2 and 3; right column). As forspeciation, extinction rate correlations improve as weconsider subsets of regimes with increasing numbers oftips, but accuracy of rate estimates is heavily influencedby the preservation rate (Fig. 4b). Preservation rates wereestimated with high accuracy, with tight distributionsaround the generating values at�= 0.01, 0.1, and 1.0 (the5 and 95% quantiles of the posterior estimate for�= 0.01are �5 =0.007 and �95 = 0.015; �5 = 0.09 and �95 = 0.12for �= 0.1; �5 = 0.96 and�95 = 1.04 for �= 1; Fig. 5).

As has been shown previously (Rabosky 2014), BAMMtends to be relatively conservative in the number ofregimes on the tree even with the high prior on thenumber of shifts used in our analyses (Fig. 6). Usingstepwise model selection with a Bayes factor thresholdof 20, we identified spurious shifts in fewer than 0.5% oftrees (Fig. 6). We tabulated the marginal posterior prob-ability of a rate shift for each branch on which a (true)rate shift occurred. These marginal probabilities variedsubstantially but generally increased with the number ofdescendent tips across all preservation regimes (Fig. 7a).The marginal probability of a shift along a particularbranch increases with both the magnitude of the rateincrease and with the number of descendant lineages(Fig. 7b). Thus, BAMM substantially underestimates thenumber of rate shifts (Fig. 6), but shifts that are notdetected tend to be those leading to small clades and/orthose that represent small changes in evolutionary rates(Fig. 7). Tip partitioning accuracy in BAMM v2.6 washigh overall, and did not vary systematically withpreservation rate (Fig. 7c).

The apparent conservatism of BAMM with respect tothe true number of shifts need not be a problem: wehave previously shown that, under the Poisson process,the vast majority of rate shifts lead to small rate regimes(<5 tips), and that shift regimes of this size are unlikelyto be inferred by any method (Rabosky et al. 2017).Because the regimes are small (e.g., contain minimaldata), their likelihood is similar across a broad range ofparameter values, and there is insufficient informationin the data to infer their existence. As pointed out byone reviewer, our results also raise a paradox: how canthe BAMM-estimated rates perform well, despite the

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 10: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 9 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 9

FIGURE 2. BAMM estimates of speciation (�), extinction (�), and relative extinction rate (�/�) estimates as a function of the true valuesfor rate regimes with at least five tips from trees fossilized with different preservation rates. Each gray point is the estimated rate for a singlerate regime. Black points represent mean estimated and inferred rates for all regimes falling within each 2% quantile of true rates. For example,the mean estimated rate of all individual regimes with a true rate between the lowest and the 2% quantile comprise the first black point. Rowsgive results for extant-only trees (1st row), low (�= 0.01; 2nd row), moderate (�= 0.1; 3rd row), and high (�= 1.0; 4th row). Trees with higherpreservation rates, and thus more fossils, have better rate estimates, with the extinction rate showing particular improvement.

massive underestimate of the number of shifts? Thesolution to this paradox lies in the fact that small rateshifts are missed by BAMM, but dominant rate classesacross the tree are captured by BAMM with reasonableaccuracy. Hence, BAMM might correctly infer only 3 of15 rate regimes in a given phylogeny, but those threeinferred regimes can govern the majority of branches inthe tree (e.g., 12 shifts leading to one or two taxa, but theother three shifts all leading to 50 or more taxa).

Sensitivity of the Method to Model ViolationsWe found that the implementation performed well

even when assumptions of homogeneous fossilizationrates through time were violated. After imposingtemporally-varying preservation rates on the data, wefound that diversification rates were still estimated withhigh accuracy (Fig. 8). The estimates of both speciationand relative extinction (Fig. 8a,b) were reasonably correl-ated with the generating values. We also found very few

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 11: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 10 1–18

10 SYSTEMATIC BIOLOGY VOL. 68

FIGURE 3. BAMM estimates of speciation (�), extinction (�), and relative extinction (�/�) rates for rate regimes with at least 20 observedtips from trees fossilized with different preservation rates (�). Each gray point represents the mean posterior estimate of rates for a regime fromthe extant-only trees (1st row), low (� = 0.01; 2nd row), moderate (�= 0.1; 3rd row), and high (�= 1.0; 4th row). Each black point representsthe value of all regimes falling within each 2% quantile (e.g., all regimes between the 0th and 2nd percent quantile for the first point). Ratesin general are better estimated in the large regimes (as compared to Fig. 2). Increased preservation rates lead to better macroevolutionary rateestimation overall, particularly for extinction.

instances where the number of inferred shifts was higherthan the generating model (Fig. 8c). Consistent with thetime-homogenous preservation simulations, 78% of thetips were accurately partitioned among the rate regimes(Fig. 8d).

Underestimating the number of fossil occurrencesinflates the speciation and extinction rates, while over-estimating the number of occurrences deflates them(Fig. 9). Individual estimates of � may be off by as

much as 50% when misspecification is exceptionallyhigh (Fig. 9a), while � may vary by substantially more(Fig. 9b). However, in aggregate across all 100 trees, themedian error at each level of occurrence misspecificationis fairly low. Encouragingly, the correlations of tree-widerate estimates for these constant rate trees are high evenwhen the fossil occurrence numbers are off by a factor often (correlation for speciation: 0.95 to 0.89; correlationfor extinction: 0.93 to 0.89). This suggests that, if the

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 12: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 11 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 11

●●

●●●●

●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●

Spec

iatio

n co

rrela

tion

0 10 20 30 40 50

0

0.25

0.5

0.75

1

●●

●●●●

●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●

●●

●●●●●●●●●●●●●●●●●●

●●●●●●●●●●●●●●●●●●●●●●●●●●●●●

●●

●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●

●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●

ψ = 1ψ = 0.1ψ = 0.01extant−only

a

Minimum regime size

●●

●●●●

●●●●●●●●●●●●

●●●●●●●●●●●●●●●●●●●

●●●●●

●●●●●●

Extin

ctio

n co

rrela

tion

0 10 20 30 40 50

0

0.25

0.5

0.75

1

●●

●●●●

●●●●●●●●●●●●

●●●●●●●●●●●●●●●●●●●

●●●●●

●●●●●●

●●

●●●●●●●●●●●●●●●●●●

●●●●●●●●●●●●●●●●●●●●●●●

●●●●●●

●●

●●●●

●●●

●●●●●●●●●●●●●

●●●●●●●●●●●●●●●●●●●●●●●●●●

●●●●

●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●

b

Minimum regime size

FIGURE 4. Speciation (�) and extinction (�) rate correlations across all regimes as a function of the minimum number of tips present ina particular subset of the data. Each point represents the correlation for speciation (a) or extinction (b) for all rate regimes with the specifiedminimum number of tips. A minimum regime size (x-axis) value of zero indicates that the correlation is computed across all rate regimes,regardless of the number of tips they contain; a minimum value of 10 would restrict this pool of regimes to only those with 10 or more tips.Minimum values of 5 and 20 correspond to results shown in Figs. 2 and 3, respectively. Larger regimes yield better rate estimates for bothspeciation and extinction. Speciation rates are reliably estimated even in the absence of fossil data, but extinction rate accuracy is markedlyimproved by the inclusion of fossils.

inferred preservation rate (ψ)

num

ber o

f tre

es

0 0.25 0.5 0.75 1 1.25

0

20

40

60

80

100

120

FIGURE 5. Marginal posterior distributions for the preservationrate (�) for trees simulated with preservation rates of 0.01, 0.1, and 1.0.Black arrows denote the true value of �, while gray bars show thedistribution of mean � across all 200 trees for each value.

number of fossil occurrences is known only with largeuncertainty, the relative rates across the phylogeny arelikely to be more accurate than the absolute values of therates (e.g., shifts and temporal heterogeneity may still bedetected reliably, even if the rate magnitude is biased).

We explored BAMM’s performance under twoscenarios where the diversification process in thesimulation model violated the assumptions of theinference model. For the diversity-dependent simula-tions, BAMM performed much better with the inclusion

of fossil information (Supplementary Fig. S11 availableon Dryad): both speciation and extinction rates wereestimated with much greater accuracy when fossils wereincluded, relative to the extant-only analysis. For themass extinction scenario, we observed consistent biasesin rate estimates when fossil and extant data werecombined, and relative extinction rates were consistentlybiased upwards (Supplementary Fig. S10 available onDryad).

Unobserved shifts seem to have a negligible impacton rate estimation. Neither the slope nor correlationcoefficient of the extinction and speciation rates forbranches within trees varies with the proportion ofunobserved rate shifts (Fig. 10), nor with the absolutenumber of unobserved shifts (see Supplementary Mater-ial available on Dryad). In contrast to the regime-specificassessments presented above, within-tree measures aredifficult to interpret under rate-shift models like BAMM.If a method does not infer rate heterogeneity withinthe tree, the metrics (especially the slope) are expectedto perform poorly. Rabosky et al. (2017) discusses thisissue in detail, documenting how the metrics may bemisleading even when rates on almost every branchare estimated with high accuracy. Regardless, thereis no obvious trend between the proportion of unob-served shifts within a tree and the overall metrics ofperformance (Fig. 10).

Empirical AnalysisOur empirical analysis of the horse phylogeny yielded

evidence for a major shift in rates leading up to the extantgenus Equus, and also that speciation and extinction ratesare approximately equal across the tree and throughtime (Fig. 11). Net diversification was found to be near-zero across the horse phylogeny (mean net div. =−0.002;

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 13: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 12 1–18

12 SYSTEMATIC BIOLOGY VOL. 68

●● ●●●●●● ● ●●●

●●●●●● ●● ●● ●● ●● ●●●

●●● ●●●●●● ●●● ● ●●●●● ● ●●●●●

●● ●●●● ●

●●●

●●●●● ●●●●●●● ●●●●●

● ●● ● ●

●●●●●●●● ●●●●●●●●

●● ●● ● ●● ●● ●● ●● ● ●● ●●●

●●●●● ●●●

● ●● ●●●● ●●● ● ●● ●●●● ●●

●●● ●●●

0 25 50 75 100 125

0

25

50

75

100

125

ψ = 0.01in

ferre

d nu

mbe

r of s

hifts

● ●●●●● ● ● ●●●

●● ●● ●●●

● ●● ●● ●● ●●●●●●●●●●

● ●●● ● ●●●●● ● ●●●●●●

●●

●●

● ●● ●●

●●

●●● ●●●●●●● ●●●●●●

●●

● ● ●●●

●●●●● ●●●●●●●●●●

●● ● ●● ●●

●●

●● ● ●● ●●●●

●●●●●

●●●

●● ●●●●●●●● ●● ●● ●● ●● ●

●●● ●●●●

0 25 50 75 100 125

ψ = 0.1

●●●● ●●

●●

●●●

● ●●●

● ●● ● ●●●●

●●●

●●●●●

● ●

●●● ●●●●● ●●●●

●●●

●●●

●●

●●

● ● ●●

●●●

● ● ●●●●●●

●●

●●●

●●●

●●●

●●●● ●●●

●●

● ●●●● ●● ● ●● ●●●

●● ●●●

●●

● ●●●●●●● ●●

● ●● ●● ●●

●● ●●●● ●●●

0 25 50 75 100 125

ψ = 1.0

● ●

●●

●●

●●

● ●

●●

●●●

●●●

●● ●

●●

●●

●●

● ●●●

●●

●●●●

●●

●●

●●

●●

●●●●●●

● ●

●●

● ●

● ●

●●

●●●

●●

● ●● ●●●

0 25 50 75 100 125

0

5

10

15

infe

rred

num

ber o

f shi

fts

● ●

●●●

● ● ●●

●●

●●

● ●

●●

●●

● ●

●●●

●● ●

●●

● ●

●●

●●

●●

●●

●●

●●●●●●

●●

●●

●● ●

●●●

●●

● ●

●●

0 25 50 75 100 125

true number of shifts

●●

●●

●●

● ●

●●●

●●●

●● ●

●●

●●

●●

●●

●●

● ●●

● ●

● ●●

● ●

● ●

0 25 50 75 100 125

FIGURE 6. The true number of shifts in the degraded trees plotted against the mean number of inferred shifts (number of non-root rateregimes best supported by stepwise model selection using Bayes factors) for preservation rates (�) of 0.01, 0.1, and 1.0. Each point represents oneof the 200 trees. The identity line is shown for reference, and the estimates slopes are very low (0.08, 0.085, and 0.18). The correlation betweentrue and estimated numbers of shifts are 0.65, 0.73, and 0.8 for the preservations rates from lowest to highest. Plots in the top row show theresults on the identical axes, while the plots in the 2nd row show the same data with the y-axis rescaled to better show the correlation betweenthe true and estimated number of shifts. Most simulated shifts are small in terms of the number of lineages and they may also reflect a relativelyminor change in rates (e.g., a shift from speciation of 0.1–0.11). As such, a method focused on summarizing the major regimes of the tree (i.e.,those with strong support) is expected to underestimate the number of shifts. This is especially true when a stringent Bayes factor threshold isused (see SOM for the mean number of shifts from the posterior distribution).

see Fig. 11), but lineage turnover rates were substantial(mean turnover = 0.7). This result is concordant withlong-standing hypotheses about the history of life, wherespeciation is only slightly elevated relative to extinction,and only in some clades (e.g., Raup 1994). However,it is worth noting that the BAMM estimates assumethat rates have been constant through time within rateregimes, and it is unlikely that this assumption is strictlytrue in practice. It is possible, for example, that thenear-zero net diversification rates estimated for equidsrepresent the long-term average of a clade that experi-enced a pronounced phase of positive net diversificationfollowed by negative net diversification rates that ledto extinction of many lineages (Quental and Marshall2013; Cantalapiedra et al. 2017). BAMM’s assumptionof time-homogeneous diversification means, in practice,that temporal rate variation will be smeared across theduration of a particular rate regime. We caution usersof BAMM from overinterpreting rates at specific pointsin time if time-constancy of rates is assumed within rateregimes.

PyRate likewise infers an increase in extinctionthrough time, but differs from the BAMM results inseveral ways. Extinction as estimated with PyRate was

substantially lower than speciation during the earliestportion of horse evolution, but speciation droppedsubstantially at roughly 15 Ma and rebounded slightlyat 6 Ma. The estimates of speciation and extinctionfrom PyRate differ substantially from those estimated byBAMM, with greater volatility in the PyRate estimatesof net diversification; BAMM inferred a more consist-ent near-zero level of net diversification. Notably, theturnover rates estimated by BAMM are much higherthan those from PyRate. Both methods estimate highpreservation rates, although as speciation, extinction,and preservation are estimated jointly, the two estimatesare different, with BAMM estimating a higher rate (�=3.69 ± 0.10) than PyRate (�= 2.27 ± 0.07).

DISCUSSION

We have described an extension of the BAMM frame-work for modeling macroevolutionary dynamics thatleverages recent advances in modeling the fossilizedbirth–death process (Stadler 2010; Ronquist et al. 2012;Heath et al. 2014; Didier et al. 2012, 2017) to estimaterates of speciation, extinction, and preservation from

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 14: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 13 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 13

1 5 50 500 5000

0

0.2

0.4

0.6

0.8

1

number of descendant taxa

mar

gina

l pro

b. o

f a s

hift

a

1 5 50 500 5000

0.01

0.1

1

10

100

1

0

marginal prob.

number of descendant taxa

chan

ge in

λ

b

0 25 50 75 100 125

0

0.2

0.4

0.6

0.8

1

●●

●●●

●●

●●

●●

●●

●●●

●●●●●

●●

●●●

●●

●●

●●●

●●●

●●●

●●●●●

●●●●●

●●●

●●

●●

●●

●●●

●●●●

●●

●●

●●●

●●●●●●●●

●●

●●

●●●●

●●

●●●

●●●

●●

●●●

●●●●●●

●●

●●●

●●●●●

●●

●●●

●●●

●●●

●●

●●

●●

●● ●

●●

●●

●●●

●●

●●●

●●

●●●

Randomψ = 0ψ = 0.01ψ = 0.1ψ = 1

number of shifts

tip−p

artit

ion

accu

racy

c

FIGURE 7. Topological accuracy of inferred shift locations. a) Marginal posterior probability of each rate shift as a function of the number oftips in the shift regime. A value of 0.50 would imply that 50% of samples from the posterior recovered a shift in the same topological location(e.g., same branch) as the true shift. In general, we expect the marginal shift probability to be somewhat less than 1.0 even for well-supportedshifts, because the evidence for a given shift can be smeared across several adjacent branches (Rabosky 2014). Red horizontal lines denote medianmarginal probabilities of rate shifts within sliding windows of arbitrary size (see text for details). b) Marginal probability of each rate shift asa function of both the proportional change in speciation rate (�NEW/�OLD) as well as the number of descendant tips in the regime. Darkerpoints represent higher marginal probabilities. Shift locations are inferred with the highest accuracy is highest when large rate increases occurand/or when shifts lead to large numbers of descendant tips (c) Accuracy with which BAMM infers cohort membership, computed as themean probability of assigning a given taxon pair to the correct grouping relationship (e.g., “same shift regime”, “different shift regime”). Redpoints represent the mean cohort accuracy of BAMM analyses across four preservation rates as a function of the number of shifts. Black squaresrepresent the corresponding cohort assignment accuracy expected under random shift placement; see text for details.

●●

● ●

●●

●●

● ●

●●

●●

●●

●●

●●

●●●●

● ●●● ●●

●●

●●

● ●

●●

●●●

●●

●●

● ●

●●

●●●●

●●●

●●●

●●●

●●

●●

●●

●●●

●●

● ●

●●●

●●●●

●● ●

●●

● ●

●●

● ●

●●

●●

●●

● ●

●●●

●● ●

●●

●●

●●

●●

● ●● ●● ●

●●●●●

●●

●●●

●●●

●●●

●● ●●●●●

● ●

●●

●●

●●●

●●

●●

●●

●●●

● ●● ●

●●

●●

●●

●●

●●

● ●

●●

●●● ●●

●●

●●●

●●

●●

●●

●●●

●●

● ●● ●

● ●●●●

● ●●●●●

● ●●

● ●

●●●●

●●

●●

●●●● ●

●●●●

● ●

●●

●●●

●●

●● ●

● ●

●● ●

●●●● ●

●●

●●

●●

● ●●

●●

●●

●●

●●

●●

●●

● ●

●●

●●

●●

●●●● ●

●●

●●

●●

● ●●●● ●

●●●

●●

●●

●●●

● ●●

●●

●●

●●●

● ●

● ●

●●

●●

●●

●●

● ●●

●●

●●

●●

● ●

●●

●●

●●

●●●

●●

●●●

●●

●●

●●●

●●

●●●

●●

●●

● ●

● ●

●●●

●●●

●●

● ●●

●●

●●

●●

●●

●●

●●

●●

●●

● ●

● ●

●●

●●

●●

●●●

●●

●●

●●

●●

●●

●●

●●

●●●● ●●

●●

●●

●● ●

●●

●●

●●

●●

● ●●

●●

●●●

●●●

● ● ●●●

●●

●●●

● ●

● ●

●●●●

●●●

● ●

●●●● ●●●

●●● ●● ●●

●●

●●● ●●●

●●

● ●

●●

●●●

● ●●●

●●●

●●

●●

●●●

●● ●●●

●●

●●

●●● ●●●

●●

●●●

●●

●●

●●●●

●●●●

●●

●●

●●

●●●

●●

●●●●●

●●

●●

●●

●●

● ●● ●●

●●

●●

● ●

●●

●●●

●●●●

●●

●●●●● ●

●●

●●

● ●

●●●●

●●●●

●●

●● ●

●● ●

●●

●●

●●●

●●

●●

●●●

● ●●

● ●●

●●●

●●

●●

●●●

●●

●●●

●●● ●

●●

●●●● ●

●●●●

●●●

●●

●●●

●● ●

●●

●●

● ●

●●●

true λ

infe

rred

λ

0.02 0.2 2

0.02

0.2

2 r = 0.85a

●●

●●●●●●

● ●

●●

●●

●●

●●●●●

●●

●●●

●●● ●●

●●●●●●

●●●●●

●● ●●●●●●●●●

●●●●●

●●

●●

●●

●●●●

●●

●●

●●●●●●●

●● ●●●

●●●●●

●●●●

●● ●●

●●●

●●●

●●

●● ●●

● ●●● ●

●●●●●●

●●●●

●●●

●●

●●●

●●●●●●

●●●

●●

●●●

●●●●●●

●●●●

●●●

●●●

●●●●●

●●●●●●●●●●

●●

●●

●●●

●●●●

●●

●●

● ●●●●●

●●

●●●●●●●●

●●

●●

●●

● ●● ●

●●●●●

●●● ●●●●

●● ●

●●●

●●●●●

●●●●

●●●● ●

● ●●●●

●● ●

●●

●●

●●

●●

●●

●●

●●●●

●●

●●●●

● ●

●●●

●●

●●

●●

● ●●●

●●

●●●●●● ●

●●●

●●●●

●●

● ●

●●● ●●●●●● ●●

●●●

●●

●●●●

●●●

●●

●●

●●

●●●

● ●●●●●●●● ●●● ●

●●●●●●●●●●●●

●●

●●

●●

●●

●●●●●

●●●

●●●

●●●●● ●

●●●●

●●

●●● ●●●●

●●●●

●●

●●●●●●

●● ●

●●●●

●●

●●

●●

●●●

●●●●●●●● ●●●

●●●

●●

● ●

●● ●●

●●●●

●●

●● ●●

●●

●●●●●●●●

●●●●●

●●

● ●●●

●●●

●●●●●

●●●●

●●

●●●

●●●● ●●●●

●●●

● ●●●●

●●●●● ●●●●● ●

● ●●●

●●

●●

●●●

●●●●●

●●●●●●●● ●●

● ●

●●●

●●

● ●

●●●

●●●

●●

●●

●●

● ●

●●●●

● ●●●

●● ●

●●●●●●●●

●●

●●●●●

● ●●●

●●

●●

●●

●●

●●

●●●●

●●

●● ●●●

●●

●● ●●

●●

●● ●●●●●

● ●●●

●● ●●●

●●

●●●●●●●●●●

●● ●●●●

●●

●●●●●●●●

●●●

●●●●

●●●●

●●●●

●●●●●

●●

●●●●● ●●

●●

●●

●●●●●● ●●●●●●●

●●●

●●● ●●

●●●

●●

●●

●●

●●

●●

●●

●●●

●●

●●

●●●●●

●●●● ● ●●●● ●● ●●●●●

●●

●●

●●●

●● ●●●●●●●●●

●●●●

● ●

●● ●●

● ●

●●●●

●●●●

●●●

● ●●●●

●●●

●●

●●●

●●

●● ● ●●

● ●●●

●●

●●●

●●●

●●●●

● ●

●●

●●● ●●

●●●●

●●●

●●●

●●

● ● ●●

●●

●●● ●●

● ●● ●●●

●●●●

●●●●●

●●●

●●●

●●●

●●

●●●

●●

●●●●●●●

●●●● ●●

true μ / λ

infe

rred

μ / λ

0 1 2 3 4

0

1

2

3

4 r = 0.65b

● ●●●●● ● ● ●●

●●● ●● ●● ●●● ●● ●●● ●● ●● ●●

●●● ●●●●●●● ●

●●●● ● ●●●●●● ●●●●●●●●

●●●

● ●●

●●

● ● ●●

●●●●● ●●●●●● ●

●● ●●

●● ●

●●●●●●●●● ●●

●●●● ●

●●● ●●

●● ●

●●●

●● ●●●

●●

●●● ● ●●● ●●●

● ●

●●●● ●●● ●●

●● ●●●●●●●●● ● ●● ●●● ●●●●●●● ●

●●● ●●●●

0 20 40 60 80

0

20

40

60

80

number of shifts

infe

rred

# of

shi

fts

c

0 20 40 60 80

0

0.2

0.4

0.6

0.8

1

●●

●●

●●●●

●●●

●●●●●

●●●●●

●●

●●●

●●●●●

●●● ●●●

●●●

●●

Randomψ ~ Uniform(0.01, 0.9)

number of shifts

tip−p

artit

ion

accu

racy

d

FIGURE 8. Performance of BAMM when the assumption of time-homogeneous fossil preservation is violated. Speciation (�; a), relativeextinction (�/�; b), inferred number of shifts (c) and cohort accuracy(d) are shown for trees fossilized using a stage-specific simulation andanalyzed using the constant-preservation rate assumption of BAMM;results illustrate regimes with five or more tips, for comparison to Fig. 2.

phylogenies that include observations of fossil (non-extant) lineages. We found that BAMM was able toinfer diversification rates, preservation rates, and thenumber and location of rate shifts with reasonableaccuracy across a range of fossilization scenarios. Theinference model implemented in BAMM is robust to atleast some violations of its assumptions, although wenecessarily explored a limited subset of many possible

ways that model assumptions can be violated. Givendebates over the inference of extinction from molecularphylogenies (Rabosky 2010, 2016; Beaulieu and O’Meara2015), it is noteworthy that inclusion of even a smallnumber of fossils (<10% of the total tips) substantiallyimproves the estimation of both the absolute and relativeextinction rates in comparison to analyses of the samedata restricted to extant tips only. BAMM’s likelihood isconditioned on the absence of unobserved rate shifts, anassumption that has been criticized in the recent literat-ure (Moore et al. 2016). However, we find no evidencethat these unknown quantities affect inference (Fig. 10).Our results are thus consistent with Rabosky et al.’s(2017) demonstration that unobserved rate shifts areunimportant for empirically relevant parameterizationsof the diversification process.

Given the preservational biases inherent to the fossilrecord, our results showing that BAMM is at leastsomewhat robust to temporal heterogeneity in preser-vation rates and uncertainty in the number of fossiloccurrences are encouraging (Figs. 8 and 9). Changesin the preservation rate through time are likely to bemore extreme in real data sets than considered by ourstudy. However, as long as the average preservationthrough time is comparable to values used in oursimulations, we expect that the inclusion of fossil datawill improve estimates of diversification parameters.Incorrect specification of the number of fossil occur-rences biases the magnitude of rate estimates, but thetrue and estimated rates remain highly correlated evenwith substantial misspecification of the true numberof fossils. This suggests that relative rates within aphylogeny may largely be robust to misspecification ofoccurrence counts, even if the absolute values of thoserates are biased. Nonetheless, assumptions about thenature of fossil preservation and lineage comparability

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 15: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 14 1–18

14 SYSTEMATIC BIOLOGY VOL. 68

●●

● ●

●●

(1) (2) (3) (4) (5) (6)

0.00

0.25

0.50

0.75

1.00

1.25

1.50

1.75

2.00

infe

rred

rate

/ tru

e ra

te

a

degree of fossil occurrence misspecification

● ●● ● ● ●

(1) (2) (3) (4) (5) (6)

0

2

4

6

8

10

12

14

16

18b

FIGURE 9. Bias in the estimation of macroevolutionary rates when the number of fossil occurrences was misspecified. The minimum numberof fossil occurrences (FMIN) is equal to the number of observed extinct tips, and the gap between the minimum and true number of fossiloccurrences (FGAP) was used to analyzed different trees at comparable levels of misspecification. The bias in speciation (a) and extinction (b)show a similar pattern of inflated rate estimates when fewer than the true number of observed fossils are specified, and deflated estimates whenthe number of fossils were exaggerated. Bold points represent the median bias, and red lines indicated the true number of fossils (vertical) andno bias in rate estimation (horizontal). The degrees of fossil misspecification are: 1) FMIN, 2) FMIN+ 0.25 × FGAP, 3) FMIN+ 0.5 × FGAP, 4) FTRUE,5) FTRUE+ FGAP, and 6) 10 × FTRUE. Trees were simulated using constant rates through time, and the inferred values are for the entire tree.

between neontological and paleontological datatypesshould be treated with caution in future empirical work.

In the case of clade-wide mass extinction(Supplementary Fig. S10 available on Dryad), the biasin BAMM’s extinction estimates is easily understood.BAMM has no ability to accommodate mass extinctionsdirectly, but it is nonetheless forced to model a situationwhere a large number of taxa became extinct ata particular point in time. Hence, the backgroundextinction rate for the focal clade must increase toaccommodate the extinction of many taxa, even thoughthe clade may have had low extinction throughoutmuch of its history. In effect, BAMM smears the effectsof the mass extinction across earlier periods of timewhen background extinction was lower.

Because the current implementation of the model doesnot accommodate time-varying diversification rates, it ispossible that resulting time-smeared rate estimates willbe an inadequate description of the true diversificationprocess. Similarly, for any clades that have gone extinctin their entirety, net diversification rates will generallybe estimated as near zero throughout the history ofthe clade. This potential for time-averaging of rates isa substantial concern for interpretation that must beconsidered carefully by users of the method, at leastuntil extensions have been developed that can betteraccommodate time-varying diversification processes.

Sampling BiasesPaleontologists have long recognized that apparent

long-term trends in macroevolutionary dynamics canresult from biases in the rock record, and in particular thetemporal variation in the total amount of sedimentaryrock (Raup 1972, 1976; Sepkoski et al. 1981; Foote and

Raup 1996; Peters and Foote 2001; Peters 2006; Holland2016). Even if total species richness is constant throughtime, intervals with greater amounts of exposed sedi-mentary rock volume, or with the same amount of rockdistributed across a greater geographic extent, will gen-erally appear to be more diverse (Raup 1972). A generaltrend of increasing preservation potential through timehas long been documented (e.g., Raup 1976), and if thistrend were universally monotonic it would be relativelyeasy to model. However, preservation rates have variedgeographically throughout Earth’s history (Peters 2005).At any given time in the past, most of the habitableenvironments that existed are unpreserved. This facetof the rock record becomes particularly problematicwhen a clade endemic to a region either appears ordisappears due to a shift in preservation potential. If� does not vary to accommodate the change in localsampling rates, then the sudden change in diversitymay be incorrectly inferred to reflect a change in thediversification parameters.

Within-lineage preservation heterogeneity can alsoimpact macroevolutionary inference, and has beendemonstrated to influence estimates of diversification(Liow et al. 2010b; Silvestro et al. 2014a,b). Factorssuch as geographic occupancy (Foote et al. 2007, 2016)and expected abundance (Stanley 1979) are thought tovary with the age of a lineage, and both can influencepreservation potential. The probability of preservationmay also vary among lineages within a phylogeny.Many factors, from body size and shell composition tohabitat preferences and foraging strategies, influencethe preservation potential of a taxon (Valentine 1989;Behrensmeyer and Hook 1992; Behrensmeyer et al. 2000;Valentine et al. 2006). Furthermore, the magnitude anddirectionality of these biases can vary with time and

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 16: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 15 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 15

= 0

reg

ress

ion

slop

e

1.0

0.5

0.0

0.5

1.0

1.5

2.0

1.0

0.5

0.0

0.5

1.0

Pea

rson

's c

orre

latio

regr

essi

on s

lope

1.0

0.5

0.0

0.5

1.0

1.5

2.0

2.5

0.0 0.2 0.4 0.6 0.8 1.0

1.0

0.5

0.0

0.5

1.0

μPea

rson

's c

orre

latio

n

= 0.01

0.0 0.2 0.4 0.6 0.8 1.0

= 0.1

0.0 0.2 0.4 0.6 0.8 1.0

proportion of shifts unobserved

= 1

0.0 0.2 0.4 0.6 0.8 1.0

FIGURE 10. Metrics of within-tree rate accuracy as a function of the proportion of shifts in the tree that are unobserved, for trees with at least50 observed tips. For the branches in each tree we evaluated the inferred and generating rate of speciation and extinction, calculated the slope(1st and 3rd rows), and Pearson correlation (2nd and 4th rows) between them, and show them relative to the proportion of rate shifts in the treethat have no sampled taxa (see Fig. 1). The relative density of values across all preservation levels is shown to the right. Both metrics for bothrates have their highest density across all trees and rate regimes near the optimal value of 1 (mean slopes of 0.67 and 0.41 for speciation andextinction; mean correlations of 0.77 and 0.48 for speciation and extinction). Neither rate shows a relationship between the within-tree metricand the proportion of unobserved shifts suggesting that unobserved shifts do not bias rate estimates in these simulations. Filled circles denotetrees where BAMM identified substantial evidence for rate variation (Bayes factor evidence >20 for one or more shifts); open circles are trees thatlacked evidence for rate variation. Note that branch-specific slopes and correlations as presented above are expected to equal zero if BAMM failsto detect diversification rate variation (see Rabosky et al. 2017, Fig. 10). For additional visualizations demonstrating the lack of a relationshipbetween performance and unobserved shifts, see the Supplementary Materials available on Dryad.

underlying sedimentology (Kidwell 2005; Foote et al.2015). The BAMM extension described in this articleincorporates a very simple model of preservation, suchthat users merely need to specify the total number offossil occurrences associated with the lineages that areincluded in the phylogeny. Accommodating changes inpreservation potential through time, among lineages,and across geographic regions is an obvious futureextension for models like BAMM that integrate extantand fossil data.

Comparability of Neontological and Paleontological Data

Many challenges remain in linking paleontologicaland neontological data, as the very concept of a lineagediffers between the two datatypes. Fossil taxa are

identified based on preserved morphological traits, suchthat that a single fossil morphospecies may be equivalentto multiple biological species that are distinguishableby multiple traits that are unlikely to fossilize (e.g.,behavior, coloration, soft-tissue anatomy, vocalizations).Further, it is difficult in the fossil record to discriminatebetween a single lineage sampled repeatedly throughtime (a series of anagenetic morphospecies) and mul-tiple, closely-related but separate lineages present indifferent stratigraphic horizons. Hence, inferences thatuse method described in this article are subject to anadditional source of uncertainty that reflects differentspecies concepts (either in theory or in practice) betweenneontological and paleontological data.

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 17: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 16 1–18

16 SYSTEMATIC BIOLOGY VOL. 68

−0.0024

0.043

0.088

0.14

net d

iv.

0.2

0.4

0.6

0.8

1.0

1.2

1.4

20 15 10 5 0

age (Ma)

rate

●●●●●●●●●●●●●●

●●●●●

●●

●●

●●

●●●●●●●●●●●●●●

BAMM μBAMM λpy μpy λ

FIGURE 11. Mean net diversification rates per branch (left), and mean speciation and extinction rates through time from BAMM and PyRate(right) for the horse phylogeny. PyRate and BAMM estimates disagree with respect to the overall turnover and net diversification rates throughtime, although estimates of speciation are in accordance at the beginning of the horse phylogeny (~20–15 Ma) and extinction estimates arecomparable toward the present (~3–0 Ma).

How Many Fossil Occurrences?

The model we have implemented assumes that fossiloccurrences are generated under a simple birth–death-fossilization process, which implies that occurrencesrepresent the sampling of a specific lineage at a spe-cific point in time. We suggest that the number ofoccurrences should reflect the number of temporally-unique occurrences of identifiable lineages, or thenumber of stratigraphically unique occurrences of bothextinct and extant species-level taxa that are includedin the phylogeny. The number of temporally-uniqueidentifiable occurrences differs from raw counts of thetotal number of fossil occurrences of the clade in severalimportant ways. In BAMM, taxonomically indeterminaterecords belonging to a clade (e.g., “Aves indet.”) shouldnot be used, as there is no way to know whether thefossil corresponds to an observed lineage within the focalclade (either an extinct or extant form represented bya lineage on the phylogeny), or if it belongs to an as-yet unrecognized branch within the clade (e.g., a newspecies not yet included in the phylogeny). Assumingwe are generally working with species-level phylogenies,one should also avoid using occurrences that have notbeen resolved below the genus level in BAMM, as theymight represent lineages that are not present in the tree.Furthermore, multiple individuals of the same speciescollected from a single bone or shell bed should betreated as a single occurrence.

As an example of how researchers might specify anappropriate number of fossil occurrences for a BAMManalysis, we will consider how occurrences might betabulated for a recent phylogenetic study of extant andextinct canid mammals (Matzke and Wright 2016). Asimple search (17 February 2017) on “Canidae” fromthe Paleobiology Database (Alroy et al. 2001; Varela etal. 2015) gave at least 10,660 fossil observations (uniquerecords). Of this set, 2946 represented stratigraphically

unique occurrences, and 2107 of those observations wereresolved to the species level. The data set analyzedby Matzke and Wright (2016) contains 142 species-leveltaxa, and we identified 1837 stratigraphically-uniqueoccurrences of those taxa in the database. For the simple(time-homogeneous) preservation model described inthis article, we would specify 1837 occurrences for aBAMM analysis of the canid phylogeny. The Paleobio-logy Database allows rapid searching for occurrencesthat match only the set of taxa present in a given phylo-geny, and querying the output for duplicate localitiescan be done quickly (see Dryad scripts for our horseexample).

At present, the model assumes that fossil occurrencesare comparable through time and across clades. Thisassumption is clearly violated in practice, particularlywhen we consider large clades whose constituent speciesdiffer in key biological traits that influence their prob-ability of preservation. To illustrate this point, considerisolated vertebrate teeth, a common type of fossil.Crocodilians shed their teeth continuously through life,although their teeth are rarely identifiable to the specieslevel. On the other hand, mammals only have two setsof teeth in their life, so they produce fewer fossils,but each tooth is far more likely to be mapped to anobserved lineage on a phylogeny. These differences makeit difficult to equate the occurrence of a single crocodiliantooth to the occurrence of a mammalian molar. Anobvious extension to the modeling framework presentedhere will be to incorporate lineage-specific variation inpreservation rates, which will help account for suchclade-specific differences in preservation potential.

CONCLUSIONS

The fossilized birth–death process has already showngreat promise for inferring and time-scaling phylogenies

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 18: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 17 1–18

2019 MITCHELL ET AL.—ESTIMATING RATES USING FOSSILS AND PHYLOGENIES 17

using all available evidence. Recent advances in extend-ing the model to incorporate piece-wise time variation(Gavryushkina et al. 2014, 2017) and non-uniform incom-plete sampling (Zhang et al. 2016) have greatly improvedthe ability to estimate phylogenies that include bothextinct and extant lineages. Herein, we have shown thatBAMM performs well on dated phylogenies that includea mix of extant and extinct tips (or extinct-only tips),providing a useful tool for characterizing diversificationrate variation across the expanding pool of time-calibrated phylogenies that include extinct lineages. Thisapproach will help researchers test hypotheses aboutmacroevolutionary rate control in clades that may lacksufficient fossil data for non-phylogenetic (occurrence-only) paleobiological inferences.

The results of this study and others (e.g., Didieret al. 2012) demonstrate that the inclusion of fossildata improves the estimation of both speciation andextinction rates from phylogenies. The assumptions ofthe model we describe are simplistic, and numerouschallenges remain to be addressed. However, we havefound that results are reasonably robust to several keyviolations of model assumptions. The method describedhere is a first step toward integrating information frompaleontological and neontological data sets to betterunderstand the extent to which the dynamics of speciesdiversification have varied through time and amonglineages.

SUPPLEMENTARY MATERIAL

Data available from the Dryad Digital Repository:https://doi.org/10.5061/dryad.50m70.

FUNDING

This work was supported by the U.S. National ScienceFoundation [NSF-DEB-1256330 to D.L.R.], by a fellow-ship from the David and Lucile Packard Foundation[D.L.R.], and by a VICI grant [R.S.E.] from the Nether-lands Organization for Scientific Research (NWO).

ACKNOWLEDGMENTS

We thank R. Ree and three anonymous reviewers forcomments on the article.

REFERENCES

Alfaro M.E., Santini F., Brock C., Alamillo H., Dornburg A., RaboskyD.L., Carnevale G., Harmon L.J. 2009. Nine exceptional radiationsplus high turnover explain species diversity in jawed vertebrates.Proc. Natl. Acad. Sci. USA. 106:13410–13414.

Alroy J. 1999. The fossil record of North American mammals: evidencefor a paleocene evolutionary radiation. Syst. Biol. 48:107–118.

Alroy J. 2008. Dynamics of origination and extinction in the marinefossil record. Proc. Natl. Acad. Sci. USA. 105:11536–11542.

Alroy, J. 2010. The shifting balance of diversity among major marineanimal groups. Science. 329: 1191–1194.

Alroy J., Marshall C.R., Bambach R.K., Bezusko K., Foote M., FürsichF.T., Hansen T.A., Holland S.M., Ivany L.C., Jablonski D., JacobsD.K., Jones D.C., Kosnik M.A., Lidgard S., Low S., Miller A.I.,Novack-Gottshall P.M., Olszewski T.D., Patzkowsky M.E., RaupD.M., Roy K., Sepkoski J.J., Sommers M.G., Wagner P.J., WebberA. 2001. Effects of sampling standardization on estimates ofPhanerozoic marine diversification. Proc. Natl. Acad. Sci. USA.98:6261–6266.

Bapst, DW. 2013. A stochastic rate-calibrated method for time-scalingphylogenies of fossil taxa. Methods Ecol. Evol. 4: 724–733.

Beaulieu J.M., O’Meara B.C. 2015. Extinction can be estimated frommoderately sized molecular phylogenies. Evolution. 69:1036–1043.

Behrensmeyer, A.K., Hook, R.W. 1992. Paleoenvironmental contextsand taphonomic modes. In: Behrensmeyer, A.K., Damuth, J.D.,DiMichele, W.A., Potts, R., Sues, H.D., Wing, S.L., editors. Terrestrialecosystems through time. Chicago: University of Chicago Press. Pp.15–136.

Behrensmeyer A.K., Kidwell S.M., Gastaldo R.A. 2000. Taphonomy andpaleobiology. Paleobiology. 26:103–147.

Benton M.J., Pearson P.N. 2001. Speciation in the fossil record. TrendsEcol. Evol. 16:405–411.

Cantalapiedra J.L., Prado J.L., Fernández M.H., Alberdi M.T. 2017.Decoupled ecomorphological evolution and diversification inNeogene-Quaternary horses. Science. 355:627–630.

Didier G., Fau M., Laurin M. 2017. Likelihood of tree topologies withfossils and diversification rate estimation. Syst. Biol. 66:964–987.

Didier G., Royer-Carenzi M., Laurin M. 2012. The reconstruc-ted evolutionary process with the fossil record. J. Theor. Biol.315:26–37.

Etienne R.S., Haegeman B. 2012. A conceptual and statistical frame-work for adaptive radiations with a key role for diversity depend-ence. Am. Nat. 180:E75–E89.

Etienne R.S., Haegeman B., Stadler T., Aze T., Pearson P.N., PurvisA., Phillimore A.B. 2012. Diversity-dependence brings molecularphylogenies closer to agreement with the fossil record. Proc. R. Soc.Lond. B Biol. Sci. 279:1300–1309.

FitzJohn R. 2010. Quantitative traits and diversification. Syst. Biol.59:619–633.

FitzJohn R.G., Maddison W.P., Otto S.P. 2009. Estimating trait-dependent speciation and extinction rates from incompletelyresolved phylogenies. Syst. Biol. 58:595–611.

Foote M., Crampton J.S., Beu A.G., Marshall B.A., Cooper R.A.,Maxwell P.A., Matcham I. 2007. Rise and fall of species occupancyin Cenozoic fossil mollusks. Science. 318:1131–1134.

Foote M., Crampton J.S., Beu A.G., Nelson C.S. 2015. Aragonite bias,and lack of bias, in the fossil record: lithological, environmental, andecological controls. Paleobiology. 41:245–265.

Foote M., Raup D.M. 1996. Fossil preservation and the stratigraphicranges of taxa. Paleobiology. 22:121–140.

Foote M., Ritterbush K.A., Miller A.I. 2016. Geographic ranges of generaand their constituent species: structure, evolutionary dynamics, andextinction resistance. Paleobiology. 42:269–288.

Gavryushkina A., Heath T.A., Ksepka D.T., Stadler T., Welch D.,Drummond A.J. 2017. Bayesian total-evidence dating reveals therecent crown radiation of penguins. Syst. Biol. 66:57–73.

Gavryushkina A., Welch D., Stadler T., Drummond A.J. 2014. Bayesianinference of sampled ancestor trees for epidemiology and fossilcalibration. PLoS Comput. Biol. 10:e1003919.

Heath T.A., Huelsenbeck J.P., Stadler T. 2014. The fossilized birth–deathprocess for coherent calibration of divergence-time estimates. Proc.Natl. Acad. Sci. USA. 111:E2957–E2966.

Holland S.M. 2016. The non-uniformity of fossil preservation. Philos.Trans. R. Soc. Lond. B Biol. Sci. 371:20150130.

Hunt G., Slater G. 2016. Integrating paleontological and phylogeneticapproaches to macroevolution. Annu. Rev. Ecol. Evol. Syst. 47:189–213.

Kidwell S.M. 2005. Shell composition has no net impact on large-scaleevolutionary patterns in mollusks. Science. 307:914–917.

Ksepka D.T., Parham J.F., Allman J.F., Benton M.J., Carrano M.T.,Cranston K.A., Donoghue P.C.J., Head J.J., Hermsen E.J., IrmisR.B., Joyce W.G., Kohli M., Lamm K.D., Leehr D., Patané J.L., Polly

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019

Page 19: University of Groningen Inferring Diversification Rate

Copyedited by: YS MANUSCRIPT CATEGORY: Regular Manuscript

[16:04 10/12/2018 Sysbio-OP-SYSB180039.tex] Page: 18 1–18

18 SYSTEMATIC BIOLOGY VOL. 68

P.D., Phillips M.J., Smith N.A., Smith N.D., Tuinen M.V., WareJ.L., Warnock R.C.M. 2015. The Fossil Calibration Database—a newresource for divergence dating. Syst. Biol. 64:853–859.

Liow L.H., Quental T.B., Marshall C.R. 2010a. When can decreasingdiversification rates be detected with molecular phylogenies andthe fossil record? Syst. Biol. 59:646–659.

Liow L.H., Skaug H.J., Ergon T., Schweder T. 2010b. Global occurrencetrajectories of microfossils: environmental volatility and the rise andfall of individual species. Paleobiology. 36:224–252.

Maddison W.P., Midford P.E., Otto S.P. 2007. Estimating a binarycharacter’s effect on speciation and extinction. Syst. Biol. 56:701–710.

Matzke, N. J., Wright, A. 2016. Inferring node dates from tip dates infossil Canidae: the importance of tree priors. Biol. Lett. 12:20160328

Mitchell J.S., Makovicky P.J. 2014. Low ecological disparity in EarlyCretaceous birds. Proc. R. Soc. Lond. B Biol. Sci. 281:20140608.

Mitchell J.S., Rabosky D.L. 2016. Bayesian model selection withBayesian analysis of macroevolutionary mixtures: effects of themodel prior on the inferred number of diversification shifts.Methods Ecol. Evol. doi: 10.1111/2041-210X.12626.

Moore, B. R., Höhna, S., May, M. R., Rannalla, B., Huelsenbeck, J. P.2016. Critically evaluating the theory and performance of Bayesiananalysis of macroevolutionary mixtures. Proc. Natl. Acad. Sci. USA.113:9569–9574.

Morlon, H. 2014. Phylogenetic approaches for studying diversification.Ecol. Lett. 17:508–525.

Morlon H., Parsons T.L., Plotkin J.B. 2011. Reconciling molecularphylogenies with the fossil record. Proc. Natl. Acad. Sci. USA.108:16327–16332.

Nee S. 2006. Birth-death models in macroevolution. Annu. Rev. Ecol.Evol. Syst. 37:1–17.

Nee S., May R.M., Harvey P.H. 1994. The reconstructed evolutionaryprocess. Philos. Trans. R. Soc. Lond. B Biol. Sci. 344:305–311.

Peters S.E. 2005. Geologic constraints on the macroevolutionary historyof marine animals. Proc. Natl. Acad. Sci. USA. 102:12326–12331.

Peters S.E. 2006. Macrostratigraphy of North America. J. Geol. 114:391–412.

Peters S.E., Foote M. 2001. Biodiversity in the Phanerozoic: a reinter-pretation. Paleobiology. 27:583–601.

Pujos F., Iuliis G.D., Cartelle C. 2016. A paleogeographic overview oftropical fossil sloths: towards an understanding of the origin ofextant suspensory sloths? J. Mamm. Evol. 1–20.

Purvis A. 2008. Phylogenetic approaches to the study of extinction.Annu. Rev. Ecol. Evol. Syst. 39:301–319.

Quental T.B. Marshall C.R. 2010. Diversity dynamics: molecu-lar phylogenies need the fossil record. Trends Ecol. Evol.25:434–441.

Quental, T.B., Marshall C. R. 2013. How the Red Queen drives terrestrialmammals to extinction. Science. 341:290–292.

Rabosky D.L. 2009. Heritability of extinction rates links diversificationpatterns in molecular phylogenies and fossils. Syst. Biol. 58:629–640.

Rabosky D.L. 2010. Extinction rates should not be estimated frommolecular phylogenies. Evolution. 64:1816–1824.

Rabosky D.L. 2014. Automatic detection of key innovations, rateshifts, and diversity-dependence on phylogenetic trees. PLoS One.9:e89543.

Rabosky D.L. 2016. Challenges in the estimation of extinction frommolecular phylogenies: a response to Beaulieu and O’Meara.Evolution. 70:218–228.

Rabosky D.L., Grundler M., Anderson C., Title P., Shi J.J., Brown J.W.,Huang H., Larson J.G. 2014. BAMMtools: an R package for theanalysis of evolutionary dynamics on phylogenetic trees. MethodsEcol. Evol. 5:701–707.

Rabosky, D. L., Mitchell, J. S., Chang J. 2017. Is BAMM flawed?Theoretical and practical concerns in the analysis of multi-ratediversification models. Syst. Biol. 66:477–498.

Rabosky D.L., Santini F., Eastman J., Smith S.A., Sidlauskas B., Chang J.,Alfaro M.E. 2013. Rates of speciation and morphological evolutionare correlated across the largest vertebrate radiation. Nat. Commun.4:1958.

Raup D.M. 1972. Taxonomic diversity during the phanerozoic. Science.177:1065–1071.

Raup D.M. 1976. Species diversity in the Phanerozoic: an interpretation.Paleobiology. 2:289–297.

Raup D.M. 1985. Mathematical models of cladogenesis. Paleobiology.11:42–52.

Raup D.M. 1994. The role of extinction in evolution. Proc. Natl. Acad.Sci. USA. 91:6758–6763.

Raup D.M., Gould S.J., Schopf T.J.M., Simberloff D.S. 1973. Stochasticmodels of phylogeny and the evolution of diversity. J. Geol. 81:525–542.

Ronquist F., Klopfstein S., Vilhelmsen L., Schulmeister S., MurrayD.L., Rasnitsyn A.P. 2012. A total-evidence approach to dating withfossils, applied to the early radiation of the hymenoptera. Syst. Biol.61:973–999.

Rosenblum E.B., Sarver B.A.J., Brown J.W., Roches S.D., HardwickK.M., Hether T.D., Eastman J.M., Pennell M.W., Harmon L.J. 2012.Goldilocks Meets Santa Rosalia: an ephemeral speciation modelexplains patterns of diversification across time scales. Evol. Biol.39:255–261.

Sakamoto, M., Benton, M. J., and C. Venditti. 2016. Dinosaurs in declinetens of millions of years before their final extinction. Proc. Natl.Acad. Sci. USA. 113:5036–5040.

Sepkoski J.J. 1978. A kinetic model of phanerozoic taxonomic diversityI. Analysis of marine orders. Paleobiology. 4:223–251.

Sepkoski J.J., Bambach R.K., Raup D.M., Valentine J.W. 1981. Phanero-zoic marine diversity and the fossil record. Nature. 293:435–437.

Sheets H.D., Mitchell C.E., Melchin M.J., Loxton J., Štorch P., CarlucciK.L., Hawkins A.D. 2016. Graptolite community responses to globalclimate change and the Late Ordovician mass extinction. Proc. Natl.Acad. Sci. USA. 113:8380–8385.

Silvestro D., Schnitzler J., Liow L.H., Antonelli A., Salamin N. 2014a.Bayesian estimation of speciation and extinction from incompletefossil occurrence data. Syst. Biol. 63:349–367.

Silvestro D., Salamin N., Schnitzler J. 2014b. PyRate: a new program toestimate speciation and extinction rates from incomplete fossil data.Methods Ecol. Evol. 5:1126–1131.

Stadler T. 2010. Sampling-through-time in birth–death trees. J. Theor.Biol. 267:396–404.

Stadler T. 2013. Recovering speciation and extinction dynamics basedon phylogenies. J. Evol. Biol. 26:1203–1219.

Stanley S.M. 1979. Macroevolution, pattern and process. Baltimore,Maryland: Johns Hopkins University Press.

Stock C. 1929. A census of the pleistocene mammals of Rancho La Brea,based on the collections of the Los Angeles Museum. J. Mammal.10:281–289.

Valentine J.W. 1989. How good was the fossil record? Clues from theCalifornian Pleistocene. Paleobiology. 15:83–94.

Valentine J.W., Jablonski D., Kidwell S., Roy K. 2006. Assessing thefidelity of the fossil record by using marine bivalves. Proc. Natl.Acad. Sci. USA. 103:6599–6604.

Varela S., González-Hernández J., Sgarbi L.F., Marshall C., Uhen M.D.,Peters S., McClennen M. 2015. paleobioDB: an R package for down-loading, visualizing and processing data from the PaleobiologyDatabase. Ecography. 38:419–425.

Zanno L.E., Makovicky P.J. 2011. Herbivorous ecomorphology andspecialization patterns in theropod dinosaur evolution. Proc. Natl.Acad. Sci. USA. 108:232–237.

Zhang C., Stadler T., Klopfstein S., Heath T.A., Ronquist F. 2016. Total-evidence dating under the fossilized birth–death process. Syst. Biol.65:228–249.

Dow

nloaded from https://academ

ic.oup.com/sysbio/article-abstract/68/1/1/4999317 by U

niversity Library user on 14 January 2019