Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-05-20T11:40:49.949Z Has data issue: false hasContentIssue false

Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction

Published online by Cambridge University Press:  16 April 2024

Bethany J. Allen*
Affiliation:
Department of Biosystems Science and Engineering, ETH Zurich, Basel, Switzerland Computational Evolution Group, Swiss Institute of Bioinformatics, Lausanne, Switzerland
Maria V. Volkova Oliveira
Affiliation:
Independent
Tanja Stadler
Affiliation:
Department of Biosystems Science and Engineering, ETH Zurich, Basel, Switzerland Computational Evolution Group, Swiss Institute of Bioinformatics, Lausanne, Switzerland
Timothy G. Vaughan
Affiliation:
Department of Biosystems Science and Engineering, ETH Zurich, Basel, Switzerland Computational Evolution Group, Swiss Institute of Bioinformatics, Lausanne, Switzerland
Rachel C. M. Warnock
Affiliation:
Geozentrum Nordbayern, Friedrich-Alexander-Universität, Erlangen, Germany
*
Corresponding author: Bethany J. Allen; Email: bethany.allen@bsse.ethz.ch
Rights & Permissions [Opens in a new window]

Abstract

Phylodynamic models can be used to estimate diversification trajectories from time-calibrated phylogenies. Here we apply two such models to phylogenies of non-avian dinosaurs, a clade whose evolutionary history has been widely debated. Although some authors have suggested that the clade experienced a decline in diversity, potentially starting millions of years before the end-Cretaceous mass extinction, others have suggested that the group remained highly diverse right up until the Cretaceous-Paleogene (K-Pg) boundary. Our results show that model assumptions, likely with respect to incomplete sampling, have a large impact on whether dinosaurs appear to have experienced a long-term decline or not. The results are also highly sensitive to the topology and branch lengths of the phylogeny used. Developing comprehensive models of sampling bias, and building larger and more accurate phylogenies, are likely to be necessary steps for us to determine whether dinosaur diversity was or was not in decline before the end-Cretaceous mass extinction.

Topics structure

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

Impact statement

Dinosaurs are well known for their abrupt demise at the end of the Cretaceous period, coincident with the Chicxulub asteroid impact at 66 Ma. However, their diversity dynamics over the course of their preceding 180-million-year history are less well understood. It is not known, for instance, whether dinosaurs were thriving or already in decline just before the impact event. This is in large part due to their highly fragmentary fossil record. Phylogenetic trees depicting evolutionary relationships provide additional information, including capturing a portion of lineage history that is otherwise not observable from fossil occurrence data. Previous analyses based on dinosaur phylogenies have reached conflicting conclusions about the evolutionary trajectory of dinosaurs before their final extinction. Here, we revisit this conflict using a phylodynamic modelling approach, which is more explicit and transparent than previous approaches, especially with respect to the assumptions made about how the dinosaur fossil record has been sampled. Using two alternative models, which differ in how they use information about the sampling process and how they model changes in the number of species through time, we show that based on available phylogenies we cannot currently reach a definitive conclusion about dinosaur diversification during the Cretaceous. More densely-sampled and accurate fossil timetrees, as well as models that capture more information about the quality of the dinosaur fossil record, may help to solve this debate.

Introduction

Dinosaurs were the dominant land animals of the Mesozoic, renowned for their diversity, disparity, and ecological novelty, but they are now represented by a single surviving subclade, birds (Brusatte et al., Reference Brusatte, Butler, Barrett, Carrano, Evans, Lloyd, Mannion, Norell, Peppe, Upchurch and Williamson2015; Benson, Reference Benson2018). The extinction of non-avian dinosaurs at the end of the Cretaceous period (approximately 66 Ma) is widely accepted to be coincident with, and likely caused by, an asteroid impact (Alvarez et al., Reference Alvarez, Alvarez, Asaro and Michel1980; Schulte et al., Reference Schulte, Alegret, Arenillas, Arz, Barton, Bown, Bralower, Christeson, Claeys, Cockell, Collins, Deutsch, Goldin, Goto, Grajales-Nishimura, Grieve, Gulick, Johnson, Kiessling, Koeberl, Kring, MacLeod, Matsui, Melosh, Montanari, Morgan, Neal, Nichols, Norris, Pierazzo, Ravizza, Rebolledo-Vieyra, Reimold, Robin, Salge, Speijer, Sweet, Urrutia-Fucugauchi, Vajda, Whale and Willumsen2010; Chiarenza et al., Reference Chiarenza, Farnsworth, Mannion, Lunt, Valdes, Morgan and Allison2020; Hull et al., Reference Hull, Bornemann, Penman, Henehan, Norris, Wilson, Blum, Alegret, Batenburg, Bown, Bralower, Cournede, Deutsch, Donner, Friedrich, Jehle, Kim, Kroon, Lippert, Loroch, Moebius, Moriya, Peppe, Ravizza, Röhl, Schueth, Sepúlveda, Sexton, Sibert, Śliwińska, Summons, Thomas, Westerhold, Whiteside, Yamaguchi and Zachos2020). However, the trajectory of dinosaur diversity throughout the Mesozoic, especially towards the end of the Cretaceous, remains controversial. A wide variety of methods have previously been used to estimate either the number of dinosaur species or their diversification rates, including interpolation or extrapolation (Fastovsky et al., Reference Fastovsky, Huang, Hsu, Martin-McNaughton, Sheehan and Weishampel2004; Wang and Dodson, Reference Wang and Dodson2006; Lloyd et al., Reference Lloyd, Davis, Pisani, Tarver, Ruta, Sakamoto, Hone, Jennings and Benton2008; Brusatte et al., Reference Brusatte, Butler, Barrett, Carrano, Evans, Lloyd, Mannion, Norell, Peppe, Upchurch and Williamson2015; Close et al., Reference Close, Evers, Alroy and Butler2018) and modelling (via regression (Barrett et al., Reference Barrett, McGowan and Page2009; Lloyd, Reference Lloyd2011; Sakamoto et al., Reference Sakamoto, Benton and Venditti2016; Bonsor et al., Reference Bonsor, Barrett, Raven and Cooper2020), species-area relationships (Russell, Reference Russell1995; Le Loeuff, Reference Le Loeuff2012), or inferring evolutionary and/or sampling rates (Starrfelt and Liow, Reference Starrfelt and Liow2016; Condamine et al., Reference Condamine, Guinot, Benton and Currie2021; Černý et al., Reference Černý, Madzia and Slater2022; Langer and Godoy, Reference Langer and Godoy2022)). Several papers have suggested that the group was already in decline before the asteroid impact (Han et al., Reference Han, Wang, Wang, Zhu, Zhou, Wang, Fang, Stidham, Wang, Wang, Li, Qin, Fan, Wen, Luo, Pan and Deng2022), and had been for the preceding 10 million years (Sloan et al., Reference Sloan, Rigby, Van Valen and Gabriel1986; Archibald, Reference Archibald2014; Condamine et al., Reference Condamine, Guinot, Benton and Currie2021), 24 million years (Sakamoto et al., Reference Sakamoto, Benton and Venditti2016), or even the whole Cretaceous (Lloyd, Reference Lloyd2011). However, others have argued that this was not the case and that dinosaurs remained highly diverse right up until the latest Cretaceous (Fastovsky et al., Reference Fastovsky, Huang, Hsu, Martin-McNaughton, Sheehan and Weishampel2004; Wang and Dodson, Reference Wang and Dodson2006; Brusatte et al., Reference Brusatte, Butler, Barrett, Carrano, Evans, Lloyd, Mannion, Norell, Peppe, Upchurch and Williamson2015; Starrfelt and Liow, Reference Starrfelt and Liow2016; Bonsor et al., Reference Bonsor, Barrett, Raven and Cooper2020). This debate sits within the context of approximately constant terrestrial tetrapod species richness throughout the Mesozoic (e.g. Benson et al., Reference Benson, Butler, Alroy, Mannion, Carrano and Lloyd2016; Close et al., Reference Close, Benson, Upchurch and Butler2017). Several possible drivers of a Cretaceous dinosaur decline have been put forward, such as environmental change resulting from Deccan Traps volcanism or sea level fluctuations, however, the poor temporal resolution of geological records at this time has hindered efforts to correlate potential causes and effects (Brusatte et al., Reference Brusatte, Butler, Barrett, Carrano, Evans, Lloyd, Mannion, Norell, Peppe, Upchurch and Williamson2015; Benson, Reference Benson2018; Chiarenza et al., Reference Chiarenza, Mannion, Lunt, Farnsworth, Jones, Kelland and Allison2019, Reference Chiarenza, Farnsworth, Mannion, Lunt, Valdes, Morgan and Allison2020). There is evidence of trophic restructuring in the latest Cretaceous, which may have left Maastrichtian food webs more vulnerable to perturbations (Mitchell et al., Reference Mitchell, Roopnarine and Angielczyk2012; García-Girón et al., Reference García-Girón, Chiarenza, Alahuhta, DeMar, Heino, Mannion, Williamson, Wilson Mantilla and Brusatte2022).

Many previous studies have commented on variability in the sampling of the dinosaur fossil record, across space, time, and clades (Wang and Dodson, Reference Wang and Dodson2006; Barrett et al., Reference Barrett, McGowan and Page2009; Benton et al., Reference Benton, Dunhill, Lloyd and Marx2011; Upchurch et al., Reference Upchurch, Mannion, Benson, Butler and Carrano2011; Benson, Reference Benson2018; Chiarenza et al., Reference Chiarenza, Mannion, Lunt, Farnsworth, Jones, Kelland and Allison2019; Cashmore et al., Reference Cashmore, Mannion, Upchurch and Butler2020; Dean et al., Reference Dean, Chiarenza and Maidment2020). This hinders efforts to accurately estimate species richness or diversification over geological time, as any true changes in diversity are likely to be obscured by sampling bias (e.g. Starrfelt and Liow, Reference Starrfelt and Liow2016; Benson et al., Reference Benson, Butler, Close, Saupe and Rabosky2021). The various methods that have been applied to estimating dinosaur diversity handle this information differently: some simply take the fossil record at face value and assume that any potential sampling biases are negligible, whereas others explicitly model the incompleteness of the fossil record and thereby infer what we do not know from the fossils we have. An example of this latter viewpoint is Chiarenza et al. (Reference Chiarenza, Mannion, Lunt, Farnsworth, Jones, Kelland and Allison2019), who used ecological niche modelling to infer where dinosaurs could have lived during the Late Cretaceous based on their environmental preferences, extrapolating beyond the area represented by known fossil occurrences.

The diversification history of a clade can be quantified using raw fossil occurrences (via approaches such as PyRate (Silvestro et al., Reference Silvestro, Salamin and Schnitzler2014; Condamine et al., Reference Condamine, Guinot, Benton and Currie2021; Černý et al., Reference Černý, Madzia and Slater2022; Langer and Godoy, Reference Langer and Godoy2022)) and phylogenetic trees (e.g. Černý et al., Reference Černý, Madzia and Slater2022; Langer and Godoy, Reference Langer and Godoy2022; Truman et al., Reference Truman, Vaughan, Gavryushkin and Gavryushkina2024). Although the fossil record contains key information about the presence of taxa at a specific place and time in the geological past, this information is highly patchy, whereas phylogenies have the advantage of capturing a portion of evolutionary history that is not directly observable (e.g. Lloyd et al., Reference Lloyd, Davis, Pisani, Tarver, Ruta, Sakamoto, Hone, Jennings and Benton2008; Starrfelt and Liow, Reference Starrfelt and Liow2016; Benson, Reference Benson2018). Tree shape, in particular the temporal distribution of node ages and branch lengths, is informative about patterns of diversification, and provides insight into parts of the tree of life which are not currently represented within the known fossil record (Lloyd et al., Reference Lloyd, Davis, Pisani, Tarver, Ruta, Sakamoto, Hone, Jennings and Benton2008).

A handful of studies have previously used phylogenies to test whether dinosaur diversity was in decline before the Cretaceous-Paleogene (K-Pg) mass extinction. One approach is using phylogenetic generalised linear mixed models (GLMMs), which assess the line of best fit between the time elapsed from the root to the tips (the predictor variable) and net speciation (the response variable). The shape and slope of this line can be used to infer whether diversity has remained constant, increased, or decreased over time. Sakamoto et al. (Reference Sakamoto, Benton and Venditti2016) applied phylogenetic GLMMs to three dinosaur supertrees and found evidence of a diversity decline, starting at least 24 million years before the end-Cretaceous mass extinction. However, subsequent discourse (Bonsor et al., Reference Bonsor, Barrett, Raven and Cooper2020; Sakamoto et al., Reference Sakamoto, Benton and Venditti2021) has raised questions about the “correct" way to apply this method, including how best to interpret mixed results and how sensitive the method is to the shape of the phylogeny. The adequacy of model fit can also be problematic (Hadfield, Reference Hadfield2010); in part, this may occur because the method fits a single smoothed curve to the entirety of the clade’s evolutionary trajectory, which does not allow for short-term fluctuations in rates to be recovered. Sakamoto et al. (Reference Sakamoto, Benton and Venditti2016) also attempted to account for sampling bias by including geological and sampling proxy data as covariates in their phylogenetic GLMMs, and found that this did not change their overall results. However, this approach does not incorporate the sampling process explicitly or formalise the relationship between diversification and sampling (Bonsor et al., Reference Bonsor, Barrett, Raven and Cooper2020; Warnock et al., Reference Warnock, Heath and Stadler2020). As a result, the effect of incomplete fossil sampling on phylogenetic GLMMs is difficult to assess.

To examine the potential impact of modelling assumptions on estimates of diversification from the non-avian dinosaur fossil record, we apply Bayesian phylodynamic models. Although phylogenetics describes the process of inferring evolutionary relationships, phylodynamics seeks to infer characteristics of the history of the clade, such as diversification rates or diversity through time (Grenfell et al., Reference Grenfell, Pybus, Gog, Wood, Daly, Mumford and Holmes2004). Here, we use two different phylodynamic models, which make different assumptions about sampling and changes in the number of species through time, to infer dinosaur diversification over the Mesozoic. Both models generate piecewise-constant trajectories, allowing parameters to be estimated within a series of predefined time intervals. The first, a coalescent model, conditions the diversification process on the observed fossil ages, treating each sample as an independent event, whereas the second, a birth-death-sampling model, instead models sampling as an explicit process that generates the observed fossil record. The number of species through time changes deterministically under the coalescent model, but under the birth-death model, this change is stochastic. In the manner by which both sampling and species numbers are treated, the coalescent model is more similar to phylogenetic GLMMs. Our results show that phylodynamic models do not conclusively support a decline in dinosaur diversity towards the end of the Cretaceous, and indicate that accurately modelling sampling bias is likely to be key to understanding diversification dynamics in deep time.

Methods

Phylogenies

We used four dinosaur supertrees, the same three as Sakamoto et al. (Reference Sakamoto, Benton and Venditti2016) in addition to a more recently constructed “metatree” (Lloyd et al., Reference Lloyd, Bapst, Friedman and Davis2017). To create a fully bifurcating topology for the metatree, we sampled 1,000 phylogenies from the set of most parsimonious trees and generated a maximum clade credibility tree using TreeAnnotator (Rambaut and Drummond, Reference Rambaut and Drummond2021).

To infer the branch lengths of the phylogenies, age range data for all non-avian dinosaur species were downloaded from the Paleobiology Database (Uhen et al., Reference Uhen, Allen, Behboudi, Clapham, Dunne, Hendy, Holroyd, Hopkins, Mannion, Novack-Gottshall, Pimiento and Wagner2023) in December 2022, with species names then matched to the tip names in the phylogenies (modifications are described in the electronic supplement). Any informal species, birds (Archaeopteryx and more bird-like species), and species without age information were removed from the phylogenies, using the ape (Paradis and Schliep, Reference Paradis and Schliep2019) and palaeoverse (Jones et al., Reference Jones, Gearty, Allen, Eichenseer, Dean, Galván, Kouvari, Godoy, Nicholl, Buffan, Flannery-Sutherland, Dillon and Chiarenza2023) packages in R (R Core Team, 2022). Following this cleaning, the smallest phylogeny contained 391 dinosaur species (Lloyd et al., Reference Lloyd, Davis, Pisani, Tarver, Ruta, Sakamoto, Hone, Jennings and Benton2008) (hereafter ‘Lloyd1’), the two medium-sized phylogenies comprised the same 542 species but differed in their topologies (Benson et al., Reference Benson, Campione, Carrano, Mannion, Sullivan, Upchurch and Evans2014) (hereafter ‘Benson1’ and ‘Benson2’), and the largest phylogeny contained 750 species (Lloyd et al., Reference Lloyd, Bapst, Friedman and Davis2017) (hereafter ‘Lloyd2’). As well as analysing the supertrees in full, we also divided each into its three major subclades (Ornithischia, Theropoda, and Sauropodomorpha). We therefore conducted our analyses on a total of 16 phylogenies.

Phylodynamic models

We used two distinct Bayesian phylodynamic models to infer diversification dynamics from these species trees: the birth-death skyline (BDSKY) model (Stadler, Reference Stadler2011; Stadler et al., Reference Stadler, Kühnert, Bonhoeffer and Drummond2013; Gavryushkina et al., Reference Gavryushkina, Welch, Stadler and Drummond2014; Heath et al., Reference Heath, Huelsenbeck and Stadler2014) and a piecewise-exponential population size model based on Kingman’s n-coalescent process (Kingman, Reference Kingman1982; Griffiths and Tavaré, Reference Griffiths and Tavaré1994). Although both models are often used in the analysis of epidemiological phylogenies, they are yet to be widely applied in macroevolution.

The BDSKY model assumes that all of the observed species are the result of a birth-death process that began with a single species at some unknown time in the past. It also assumes that time is divided into one or more intervals; here, we defined eight time bins based on different geological intervals (see below). Within a single interval i, species give rise to new species at the constant rate λi (per co-existing species per Myr), and go extinct at a constant rate μi (per species per Myr). Additionally, fossils are produced at rate ѱi (per co-existing species per Myr). Species are not removed after sampling, allowing (in principle) sampled species to be direct ancestors of one another (Gavryushkina et al., Reference Gavryushkina, Welch, Stadler and Drummond2014). As our phylogenies only include non-avian (extinct) dinosaurs, we assume no extant sampling ( = 0), and condition the model on producing at least one fossil. A diversification (λi - μi) rate was calculated post-hoc for each interval in each iteration.

The piecewise-exponential coalescent model we use assumes that the observed tree is the result of a coalescent process parameterised by a time-dependent effective population size function, N(t). At any given time, the value of this function can be interpreted as proportional to a number of extant species, and thus we also refer to it as the effective species richness. We assume that this function has a continuous, piecewise-exponential form, with growth rates in each interval given by the diversification rate parameter, ri, together with the effective species richness at the end of the most recent interval (here, the Coniacian–Maastrichtian), N f. A key difference between this model and the BDSKY model is that the coalescent does not explicitly model the sampling process; it simply assumes that the sample dates (fossil ages) are independent of the number of species over time, and that the species sampled are drawn randomly from all co-existing species.

We used the boundaries of eight geological intervals of approximately equal length (Early–Mid Triassic, prior to 237 Ma; Late Triassic, 201.4–237.0 Ma; Early Jurassic, 174.7–201.4 Ma; Middle Jurassic, 161.5–174.7 Ma; Late Jurassic, 145.0–161.5 Ma; Berriasian–Barremian, 121.4–145.0 Ma; Aptian–Turonian, 89.8–121.4 Ma; Coniacian–Maastrichtian, 66.0–89.8 Ma (Cohen et al., Reference Cohen, Finney, Gibbard and Fan2013)) as the change times for the piecewise rates in all models.

Bayesian inference of model parameters

Our models constitute a specific hypothesis for how the empirical phylogenetic tree Τ was produced, and are evaluated using the probability of observing this tree given the model-specific parameters. We used Bayesian inference to infer these model parameters, as well as the branch lengths, from the predetermined phylogenetic relationships in the supertrees and the imposed tip constraints.

Specifically, conditional on a phylogenetic tree Τ and a particular phylodynamic model M, we seek to infer the model parameters Θ M. In the case that M is the BDSKY model, Θ BDSKY = $ \left\{\overrightarrow{\lambda},\overrightarrow{\mu},\overrightarrow{\unicode{x03C8}},{t}_{or}\right\} $ , whereas when M is the coalescent model Θ C = $ \left\{\overrightarrow{N},\overrightarrow{r}\right\} $ . In the Bayesian context, inference amounts to characterisation of the posterior distribution

$$ P\left({\Theta}_M|T,M\right)=\frac{P\left(T|M,{\Theta}_M\right)P\left({\Theta}_M|M\right)}{P\left(T|M\right)}, $$

where P(T|M M) is the likelihood of the model parameters given the tree under the particular model M, P M|M) is the prior probability distribution for the model parameters, and P(T|M) is the marginal likelihood of the model (which is constant with respect to the model parameters). In both the BDSKY and coalescent models, we express P M|M) as a product of priors for each of the individual model parameters, meaning that we assume no correlation between these individual parameters.

The prior probability distributions used for the individual parameters are listed in Table 1. The scale of the birth, death, and diversification rate priors was based on estimates from a study calculating diversification rates in a large number of extant and extinct phylogenies (0.02 to 1.54 speciation/extinction events per lineage per million years) (Henao Diaz et al., Reference Henao Diaz, Harmon, Sugawara, Miller and Pennell2019). In the absence of robust methods for estimating sampling completeness from the fossil record, our prior on the sampling rate favours smaller values (the mean represents one sample per lineage per 5 million years; Table 1) but does not explicitly exclude larger values.

Table 1. Priors for the Bayesian phylodynamic analyses

The branch lengths of the phylogenies were also inferred within the BDSKY and coalescent analyses. Tip constraints were placed on each species, in the form of a uniform probability distribution ranging from the oldest possible age of the oldest fossil to the youngest possible age of the youngest fossil. In each MCMC iteration, the age of each fossil, together with internal node ages and phylodynamic parameters, was sampled. Through this joint inference, we take into account uncertainty in the branch lengths, with the origination and extinction times of each lineage occurring before and after the sampled fossil age, respectively. In this analysis, the K-Pg boundary is treated as analogous to the “present day” when analysing extant phylogenies, allowing branches to reach the boundary without becoming extinct. Modelling sampling and extinction processes separately makes our approach robust to issues such as the Signor-Lipps effect (Signor and Lipps, Reference Signor and Lipps1982).

Both of the models described were implemented using the phylogenetic inference software BEAST 2 (Barido-Sottani et al., Reference Barido-Sottani, Bošková, du Plessis, Kühnert, Magnus, Mitov, Müller, PečErska, Rasmussen, Zhang, Drummond, Heath, Pybus, Vaughan and Stadler2018; Bouckaert et al., Reference Bouckaert, Vaughan, Barido-Sottani, Duchêne, Fourment, Gavryushkina, Heled, Jones, Kühnert, De Maio, Matschiner, Mendes, Müller, Ogilvie, du Plessis, Popinga, Rambaut, Rasmussen, Siveroni, Suchard, Wu, Xie, Zhang, Stadler and Drummond2019), using its MCMC algorithm to sample the above posterior probability distributions conditional on each of the trees. Each BDSKY and coalescent chain was run until the effective sample size for each model parameter was greater than 200, and therefore considered to have converged. The first 10% of each chain was discarded as burn-in prior to further analysis. Subsequent data processing and figure plotting was carried out in R (R Core Team, 2022). All relevant BEAST 2 input files and R scripts are available in the electronic supplement.

Results

The results of the coalescent analyses conducted using the 16 phylogenies are summarised in Figure 1, with the corresponding numerical estimates shown in the Supplementary Tables. There is clear variation in the median diversification estimates obtained, and the width of the error bars, between the analyses based on different clades, phylogenies, and branch lengths. However, most of the exponential coalescent models indicate a small but negative diversification rate in dinosaurs in the Coniacian–Maastrichtian (69.8% of posterior negative for Lloyd1, 96.5% for Benson1, 96.8% for Benson2, 94.1% for Lloyd2), and for some, this is also true of the preceding Aptian–Turonian (99.2% of posterior negative for Lloyd1, 68.2% for Benson1, 68.9% for Benson2, 100.0% for Lloyd2). Positive diversification rates are generally favoured in all other time bins, with the exception of the Early Jurassic (94.4% of posterior negative for Lloyd1, 98.9% for Benson1, 89.7% for Benson2, 99.7% for Lloyd2). In the full phylogenies and all three subclades, diversification rate uncertainty is highest in the Early–Mid Triassic and tends to decrease over the Mesozoic. There is most disagreement between the phylogenies for the sauropods, with the smallest (Lloyd1) phylogeny showing opposite diversification trends to the other three. However, for all of the phylogenies, most of the posterior probability lies on a strong Coniacian–Maastrichtian decline for the clade (100.0% of posterior negative for Lloyd1, 90.2% for Benson1, 90.2% for Benson2, 100.0% for Lloyd2). For theropods, a small but negative diversification rate is inferred immediately before the K-Pg boundary (99.4% of posterior negative for Lloyd1, 99.9% for Benson1, 99.9% for Benson2, 97.0% for Lloyd2). For the ornithischians, the two Benson and larger Lloyd trees all indicate a latest Cretaceous decline (94.5% of posterior negative for Benson1, 94.6% for Benson2, 97.3% for Lloyd2), whereas the smaller Lloyd tree suggests no substantial change in diversity during this interval (49.2% of posterior negative).

Figure 1. Diversification rates estimated using the piecewise-exponential coalescent model. Time moves forwards from left to right along the x-axis, with the K-Pg boundary at the end of the Coniacian–Maastrichtian bin. Estimates are shown for each of four phylogenies, ordered from smallest to largest. Points show the median values, and error bars indicate 95% highest posterior density. Dinosaur silhouettes for Ornithischia (top right), Sauropodomorpha (bottom left) and Theropoda (bottom right) are from Phylopic.

The results of the birth-death analyses are summarised in Figures 2 and 3, which show the piecewise-constant estimates of diversification and sampling rates, respectively, from each of the phylogenies, and Supplementary Tables, which provide the estimated parameter values. There is less variation in the BDSKY results between the different subclades, and based on the different tree topologies, than in the coalescent results.

Figure 2. Diversification rates estimated using the piecewise-constant fossilised birth-death skyline model. Time moves forwards from left to right along the x-axis, with the K-Pg boundary at the end of the Coniacian–Maastrichtian bin. Estimates are shown for each of four phylogenies, ordered from smallest to largest. Points show the median values, and error bars indicate 95% highest posterior density. Dinosaur silhouettes for Ornithischia (top right), Sauropodomorpha (bottom left) and Theropoda (bottom right) are from Phylopic.

Figure 3. Sampling rates estimated using the piecewise-constant fossilised birth-death model. Time moves forwards from left to right along the x-axis, with the K-Pg boundary at the end of the Coniacian–Maastrichtian bin. Estimates are shown for each of four phylogenies, ordered from smallest to largest. Points show the median values, and error bars indicate 95% highest posterior density. Dinosaur silhouettes for Ornithischia (top right), Sauropodomorpha (bottom left) and Theropoda (bottom right) are from Phylopic.

The most apparent pattern is that all of the models have much greater uncertainty on diversification rates in the final time bin, the Coniacian–Maastrichtian (Figure 2). This is coupled with an increase in the inferred sampling rates during this interval (Figure 3). In the full phylogenies and all three subclades, the scale of this effect decreases with increasing phylogeny size.

Despite this, in the BDSKY analyses, all four phylogenies place most posterior probability on a positive diversification rate for dinosaurs in the latest Cretaceous (90.0% of posterior positive for Lloyd1, 98.0% for Benson1, 98.1% for Benson2, 99.9% for Lloyd2). In all three subclades, it is more unclear as to whether diversification was positive or negative, or simply constant, prior to the K-Pg boundary. All of the models appear to favour positive diversification in the Late Jurassic (99.9% of posterior positive for Lloyd1, 100.0% for Benson1, 100.0% for Benson2, 100.0% for Lloyd2), and also in the Aptian–Turonian (98.4% of posterior positive for Lloyd1, 98.4% for Benson1, 97.4% for Benson2, 99.2% for Lloyd2).

Discussion

In this study, we characterise dinosaur diversification using two different phylodynamic models: the birth-death-sampling (BDSKY) and coalescent skyline models. The coalescent model recovered a downturn in diversity during the latest Cretaceous with a posterior probability of 97% using the Benson phylogenies, and a posterior probability of 94% using the larger Lloyd phylogeny (Figure 1). The BDSKY model instead inferred an increase in dinosaur diversity in the latest Cretaceous with a posterior probability of more than 98% based on these three largest phylogenies (Figure 2). Our results therefore span the range of diversification estimates obtained using other methods in previous literature. The difference in results we obtained using the two phylodynamic models can be linked directly to the different assumptions they make, highlighting that modelling decisions, whether conscious or unconscious, can qualitatively impact estimated diversification trajectories.

First, although the number of species through time changes stochastically in the birth-death model, this change is a deterministic (exponential) function of the parameters in the coalescent model. This contrast should have the largest impact when the number of species is very low, meaning both early in the history of the clade, and just prior to total extinction. As a result, we might expect to see the greatest difference between the model results in the first and last time bins (so this could be considered an “edge effect”). This effect may be contributing to the stark difference in our diversification estimates for the Coniacian–Maastrichtian time bin between the two models.

Second, the coalescent model assumes no relationship between species richness dynamics and the number and times of the samples: practically, each fossilisation event is treated as an independent phenomenon. The birth-death model instead treats sampling as a process, parameterised in the model as a rate (which is constant within each time bin). This rate is dependent upon the number of lineages, and therefore species, present at that time. The Coniacian–Maastrichtian is the most heavily sampled interval in our dataset (Supplementary Figure 1; Close et al., Reference Close, Benson, Upchurch and Butler2017; Reference Close, Benson, Alroy, Behrensmeyer, Benito, Carrano, Clearly, Dunne, Mannion, Uhen and Butler2019), and for the coalescent analyses, this results in relatively narrow HPD intervals on these diversification estimates in comparison with the other time bins (Figure 1). In contrast, in all of the birth-death-sampling analyses, we see drastically elevated uncertainty in estimated diversification rates for the Coniacian–Maastrichtian (Figure 2), with corresponding high uncertainty in the sampling rate (Figure 3). The birth-death-sampling model cannot discern whether this increased density of fossil sampling is due to a higher sampling or diversification rate, as reflected in the posterior distributions. However, we also see that the width of the HPD intervals for estimated diversification and sampling rates in the Coniacian–Maastrichtian decrease with increasing phylogeny size (Figures 2 and 3). Providing the birth-death-sampling model with more data therefore seems to reduce the uncertainty in our parameter estimates; increasing the size of the phylogenies used to conduct our skyline analyses may therefore allow us to infer more accurate diversification estimates in future.

The birth and death rates estimated in our birth-death-sampling models are, in some cases, fairly high in relation to previous estimates: median values for some phylogenies reach more than three events per lineage per million years (Supplementary Table 1), whereas Henao Diaz et al. (Reference Henao Diaz, Harmon, Sugawara, Miller and Pennell2019) estimated 0.02 to 1.54 events across a variety of clades, and Lloyd et al. (Reference Lloyd, Bapst, Friedman and Davis2017) estimated 0.94 events for dinosaurs. However, these models generally estimated relatively low diversification rates (Figure 2), with birth and death rates closely coupled in all time intervals, except the Coniacian-Maastrichtian, for all phylogenies (Supplementary Table 1). This coupling has previously been observed in other analyses of diversification in the fossil record (Alroy, Reference Alroy2008; Henao Diaz et al., Reference Henao Diaz, Harmon, Sugawara, Miller and Pennell2019; Černý et al., Reference Černý, Madzia and Slater2022), and suggests that while diversification can probably be estimated fairly reliably, disentangling speciation and extinction rates is more difficult.

Although the birth-death-sampling results suggest that all three dinosaur clades maintained their diversity or underwent slightly positive diversification throughout the Cretaceous, the coalescent results suggest that dinosaur diversity may have been in decline from the Aptian to Maastrichtian (Figures 1 and 2). The coalescent models suggest that although ornithischians and theropods may only have experienced a Coniacian-Maastrichtian decline, sauropodomorphs may have had negative diversification rates between the Aptian and Maastrichtian. This is consistent with other previous studies which found that ornithischians may have had higher diversification rates in the Cretaceous, particularly hadrosaurs and ceratopsids (Lloyd et al., Reference Lloyd, Davis, Pisani, Tarver, Ruta, Sakamoto, Hone, Jennings and Benton2008; Barrett et al., Reference Barrett, McGowan and Page2009; Sakamoto et al., Reference Sakamoto, Benton and Venditti2016), alongside a previously reported reduction in the number of sauropodomorph fossils into the Late Cretaceous (Barrett and Upchurch, Reference Barrett, Upchurch, Curry-Rogers and Wilson2005; Mannion et al., Reference Mannion, Upchurch, Carrano and Barrett2011; Starrfelt and Liow, Reference Starrfelt and Liow2016). Positive Late Jurassic diversification rates in all clades suggested by the birth-death models correspond to an observed peak in local dinosaur richness (Close et al., Reference Close, Benson, Alroy, Behrensmeyer, Benito, Carrano, Clearly, Dunne, Mannion, Uhen and Butler2019).

Previous attempts to include sampling bias in methods that estimate diversification have used proxy data, measurable variables thought to correlate with less tangible factors affecting diversity in the fossil record. The number of dinosaur-bearing geological formations is a commonly used example, thought to correlate with the amount of terrestrial rock outcrop area for each geological stage, which is expected to be a strong influence on the age distribution of collected fossils (Wang and Dodson, Reference Wang and Dodson2006; Barrett et al., Reference Barrett, McGowan and Page2009; Lloyd, Reference Lloyd2011; Upchurch et al., Reference Upchurch, Mannion, Benson, Butler and Carrano2011; Starrfelt and Liow, Reference Starrfelt and Liow2016). Sakamoto et al. (Reference Sakamoto, Benton and Venditti2016) used proxy data as a covariate in their phylogenetic GLMMs, and some modelling approaches have used various types of proxy data to try and extract “residual" patterns of dinosaur diversity (Barrett et al., Reference Barrett, McGowan and Page2009; Lloyd, Reference Lloyd2011). However, simulation studies have demonstrated that residual modelling, particularly using geological proxies, may degrade the biological signal in the data rather than eliminating bias (Brocklehurst, Reference Brocklehurst2015; Sakamoto et al., Reference Sakamoto, Venditti and Benton2017; Dunhill et al., Reference Dunhill, Hannisdal, Brocklehurst and Benton2018). A proxy-based approach also fails to acknowledge the wide variety of biases that affect the fossil record (Raup, Reference Raup1976), such as Lagerstätten effects (Walker et al., Reference Walker, Dunhill and Benton2020), preservation biases based on morphology (Brusatte et al., Reference Brusatte, Butler, Barrett, Carrano, Evans, Lloyd, Mannion, Norell, Peppe, Upchurch and Williamson2015; Benson, Reference Benson2018), and “dark data" in museums and private collections (Marshall et al., Reference Marshall, Finnegan, Clites, Holroyd, Bonuso, Cortez, Davis, Dietl, Druckenmiller, Eng, Garcia, Estes-Smargiassi, Hendy, Hollis, Little, Nesbitt, Roopnarine, Skibinski, Vendetti and White2018). Aside from rock outcrop area, the geography of fossil collection is also greatly driven by political and socio-economic factors (Raja et al., Reference Raja, Dunne, Matiwane, Khan, Nätscher, Ghilardi and Chattopadhyay2022). For dinosaurs specifically, the known record is highly concentrated in North America (Hurlbert and Archibald, Reference Hurlbert and Archibald1995; Le Loeuff, Reference Le Loeuff2012; Brusatte et al., Reference Brusatte, Butler, Barrett, Carrano, Evans, Lloyd, Mannion, Norell, Peppe, Upchurch and Williamson2015; Chiarenza et al., Reference Chiarenza, Mannion, Lunt, Farnsworth, Jones, Kelland and Allison2019; Dean et al., Reference Dean, Chiarenza and Maidment2020; Han et al., Reference Han, Wang, Wang, Zhu, Zhou, Wang, Fang, Stidham, Wang, Wang, Li, Qin, Fan, Wen, Luo, Pan and Deng2022), although the strength of this bias has reduced, and increasingly been accounted for, over time (e.g. Close et al., Reference Close, Benson, Alroy, Behrensmeyer, Benito, Carrano, Clearly, Dunne, Mannion, Uhen and Butler2019). Fossil abundance metrics have also been used, but even these are an imperfect proxy for sampling bias, especially when integrating data from phylogenies and fossil databases that do not contain the same taxa.

The approaches to sampling used by our coalescent and birth-death-sampling models are also not a perfect fit for the true nature of the fossil record, and violations of both models’ sampling assumptions may be biasing our results (e.g. Karcher et al., Reference Karcher, Palacios, Bedford, Suchard and Minin2016). However, methods for estimating diversity which attempt to mechanistically model sampling in a more realistic way will likely be a necessary step in unravelling how fossil record bias impacts our understanding of biodiversity in deep time (Brusatte et al., Reference Brusatte, Butler, Barrett, Carrano, Evans, Lloyd, Mannion, Norell, Peppe, Upchurch and Williamson2015; Starrfelt and Liow, Reference Starrfelt and Liow2016; Černý et al., Reference Černý, Madzia and Slater2022). Aside from this, there are additional ways in which the approach we used might be improved in future. Both of our models expect that sampling is randomly distributed across co-existing lineages, an assumption held by most approaches to estimating diversity in the fossil record, but which is not true (Hurlbert and Archibald, Reference Hurlbert and Archibald1995; Wang and Dodson, Reference Wang and Dodson2006; Benson, Reference Benson2018; Černý et al., Reference Černý, Madzia and Slater2022). Multi-type models may be used to allocate species to categories with different sampling parameters (Kühnert et al., Reference Kühnert, Stadler, Vaughan and Drummond2016), however more thought is needed on how best to assign species to discrete categories. Piecewise-constant models, as used in this paper, may be vulnerable to inaccuracies when large fluctuations in rates are present within a single bin (similarly to TRiPS (Starrfelt and Liow, Reference Starrfelt and Liow2016)), and more understanding of how change time choice is important for achieving convergence and obtaining meaningful rate estimates is needed (e.g. Allen et al., Reference Allen, Vaughan, du Plessis, Schouten, Yuan, Willett and Stadler2024). Others have also commented on the sensitivity of models to input parameters and priors more broadly (Starrfelt and Liow, Reference Starrfelt and Liow2016; O’Reilly and Donoghue, Reference O’Reilly and Donoghue2020; Černý et al., Reference Černý, Madzia and Slater2022), and highlighted the importance of making careful, informed decisions when choosing analyses and carrying out model adequacy tests when possible. Careful prior choice is also required to avoid rate non-identifiability (Smiley, Reference Smiley2018; Louca and Pennell, Reference Louca and Pennell2020; Černý et al., Reference Černý, Madzia and Slater2022), although piecewise-constant methods may be more robust to this problem than those which generate continuous curves (Legried and Terhorst, Reference Legried and Terhorst2022; Truman et al., Reference Truman, Vaughan, Gavryushkin and Gavryushkina2024).

Previous authors have commented on the necessity of continuing to collect new fossils to improve our knowledge of dinosaur evolutionary dynamics (e.g. Benson, Reference Benson2018; Bonsor et al., Reference Bonsor, Barrett, Raven and Cooper2020; Černý and Simonoff, Reference Černý and Simonoff2023), to which we would add that there are also many ways in which we could make better use of the fossils and data we already have. With further model development, full Bayesian phylodynamic inference of the tree and model parameters may become possible, allowing estimation of evolutionary rates across uncertainty in the topology and branch lengths of the phylogeny. This would address issues around the sensitivity of results to tree shape in currently available methods (shown here but also by Bonsor et al. (Reference Bonsor, Barrett, Raven and Cooper2020) and Sakamoto et al. (Reference Sakamoto, Benton and Venditti2021)). Such an approach could also allow for the inclusion of more data, such as the incorporation of more fossil age information (Stadler et al., Reference Stadler, Gavryushkina, Warnock, Drummond and Heath2018; Warnock et al., Reference Warnock, Heath and Stadler2020), and utilising fossils both with and without character data (Andréoletti et al., Reference Andréoletti, Zwaans, Warnock, Aguirre-Fernández, Barido-Sottani, Gupta, Stadler and Manceau2022). Larger phylogenies may also enable such a model to infer evolutionary rates at a finer temporal resolution. Between the results presented here and the aforementioned potential for future improvement, it is clear that phylodynamic models can provide important insights into macroevolutionary processes.

Conclusions

The trajectory of non-avian dinosaur diversification prior to their demise at the K-Pg boundary has been fiercely debated. Here, we apply two phylodynamic models to dinosaur phylogenies, to investigate the influence of sampling assumptions on estimates of evolutionary rates. Our birth-death-sampling skyline model results do not support a Cretaceous downturn in dinosaur diversity, whereas the piecewise-exponential coalescent model results do. This disparity in results indicates that fundamental differences in model design, especially with respect to sampling, can have a dramatic influence on estimates of diversification. It also highlights the importance of understanding model assumptions more broadly, providing context for results and facilitating comparison between models. Future work examining the fit of existing phylodynamic models to palaeontological datasets will help to illuminate whether one model should be favoured above the other and highlight areas for future model development.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/ext.2024.5.

Supplementary material

The supplementary material for this article can be found at http://doi.org/10.1017/ext.2024.5.

Data availability statement

All files and code necessary to run these analyses are available in the associated Zenodo repository https://doi.org/10.5281/zenodo.10996357. A Taming the Beast tutorial (Barido-Sottani et al., Reference Barido-Sottani, Bošková, du Plessis, Kühnert, Magnus, Mitov, Müller, PečErska, Rasmussen, Zhang, Drummond, Heath, Pybus, Vaughan and Stadler2018) explaining how to apply these methods (using these analyses as a case study) is also available at https://taming-the-beast.org/tutorials/.

Acknowledgements

We thank members of the Computational Evolution group, particularly Louis du Plessis, for helpful discussion, and Graeme Lloyd for assistance with using the phylogenies. We thank the anonymous reviewers for their comments, which helped to improve the manuscript. All BEAST 2 analyses were conducted using the ETH Zurich Euler cluster. We thank those who have contributed to the Paleobiology Database data used in this study. This is Paleobiology Database Official Publication No. 481. We also thank the artists who contributed the Phylopic images used in our figures: Matt Dempsey, Jaime Headden, and Iain Reid.

Author contribution

R.C.M.W., T.S. and T.G.V. developed the study concept. M.V.V.O. and B.J.A. wrote the code, with help from R.C.M.W. and T.G.V. B.J.A. and M.V.V.O. conducted the analyses, and B.J.A. made the figures. B.J.A., R.C.M.W. and T.G.V. wrote the manuscript. All authors edited the manuscript and approved its final version.

Financial support

This work was supported by the Amgen Scholars Programme (MVVO) and an ETH+ grant (BJA, “BECCY").

Competing interest

The authors declare no competing interest exist.

Footnotes

Tanja Stadler, Timothy G. Vaughan and Rachel C. M. Warnock contributed equally to this work.

References

Allen, BJ, Vaughan, TG, du Plessis, L, Schouten, TLA, Yuan, Z, Willett, SD and Stadler, T (2024) Inferring the evolutionary history of the Sino-Himalayan biodiversity hotspot using a Bayesian birth-death skyline model. Geological Society London Special Publications 549, in press.CrossRefGoogle Scholar
Alroy, J (2008) Dynamics of origination and extinction in the marine fossil record. Proceedings of the National Academy of Sciences USA 105, 1153611542.CrossRefGoogle ScholarPubMed
Alvarez, LW, Alvarez, W, Asaro, F and Michel, HV (1980) Extraterrestrial cause for the Cretaceous-Tertiary extinction. Science 208, 10951108.CrossRefGoogle ScholarPubMed
Andréoletti, J, Zwaans, A, Warnock, RCM, Aguirre-Fernández, G, Barido-Sottani, J, Gupta, A, Stadler, T and Manceau, M (2022) The occurrence birth–death process for combined-evidence analysis in macroevolution and epidemiology. Systematic Biology 71, 14401452.CrossRefGoogle ScholarPubMed
Archibald, JD (2014) What the dinosaur record says about extinction scenarios. Geological Society of America Special Papers 505, 213224.Google Scholar
Barido-Sottani, J, Bošková, V, du Plessis, L, Kühnert, D, Magnus, C, Mitov, V, Müller, NF, PečErska, J, Rasmussen, DA, Zhang, C, Drummond, AJ, Heath, TA, Pybus, OG, Vaughan, TG and Stadler, T (2018) Taming the BEAST – A community teaching material resource for BEAST 2. Systematic Biology 67, 170174.CrossRefGoogle ScholarPubMed
Barrett, PM, McGowan, AJ and Page, V (2009) Dinosaur diversity and the rock record. Proceedings of the Royal Society B 276, 26672674.CrossRefGoogle ScholarPubMed
Barrett, PM and Upchurch, P (2005) Sauropod diversity through time: Possible macroevolutionary and paleoecological implications. In Curry-Rogers, KA and Wilson, JA (eds.), The Sauropods: Evolution and Paleobiology. Berkeley CA: University of California Press, pp. 125156.Google Scholar
Benson, RBJ (2018) Dinosaur macroevolution and macroecology. Annual Review of Ecology, Evolution, and Systematics 49, 379408.CrossRefGoogle Scholar
Benson, RBJ, Butler, RJ, Alroy, J, Mannion, PD, Carrano, MT and Lloyd, GT (2016) Near stasis in the long-term diversification of Mesozoic tetrapods. PLoS Biology 14, e1002359.CrossRefGoogle ScholarPubMed
Benson, RBJ, Butler, RJ, Close, RA, Saupe, EE and Rabosky, DL (2021) Biodiversity across space and time in the fossil record. Current Biology 31, R1225R1236.CrossRefGoogle ScholarPubMed
Benson, RBJ, Campione, NE, Carrano, MT, Mannion, PD, Sullivan, C, Upchurch, P and Evans, DC (2014) Rates of dinosaur body mass evolution indicate 170 million years of sustained ecological innovation on the avian stem lineage. PLoS Biology 12, e1001853.CrossRefGoogle ScholarPubMed
Benton, MJ, Dunhill, AM, Lloyd, GT and Marx, FG (2011) Assessing the quality of the fossil record: Insights from vertebrates. Geological Society London Special Publications 358, 6394.CrossRefGoogle Scholar
Bonsor, JA, Barrett, PM, Raven, TJ and Cooper, N (2020) Dinosaur diversification rates were not in decline prior to the K-Pg boundary. Royal Society Open Science 7, 201195.CrossRefGoogle ScholarPubMed
Bouckaert, R, Vaughan, TG, Barido-Sottani, J, Duchêne, S, Fourment, M, Gavryushkina, A, Heled, J, Jones, G, Kühnert, D, De Maio, N, Matschiner, M, Mendes, FK, Müller, NF, Ogilvie, HA, du Plessis, L, Popinga, A, Rambaut, A, Rasmussen, D, Siveroni, I, Suchard, MA, Wu, C-H, Xie, D, Zhang, C, Stadler, T and Drummond, AJ (2019) BEAST 2.5: An advanced software platform for Bayesian evolutionary analysis. PLoS Computational Biology 15, e1006650.CrossRefGoogle ScholarPubMed
Brocklehurst, N (2015) A simulation-based examination of residual diversity estimates as a method of correcting for sampling bias. Palaeontologia Electronica 18, 115.Google Scholar
Brusatte, SL, Butler, RJ, Barrett, PM, Carrano, MT, Evans, DC, Lloyd, GT, Mannion, PD, Norell, MA, Peppe, DJ, Upchurch, P and Williamson, TE (2015) The extinction of the dinosaurs. Biological Reviews 90, 628642.CrossRefGoogle ScholarPubMed
Cashmore, DD, Mannion, PD, Upchurch, P and Butler, RJ (2020) Ten more years of discovery: Revisiting the quality of the sauropodomorph dinosaur fossil record. Palaeontology 63, 951978.CrossRefGoogle Scholar
Černý, D, Madzia, D and Slater, GJ (2022) Empirical and methodological challenges to the model-based inference of diversification rates in extinct clades. Systematic Biology 71, 153171.CrossRefGoogle Scholar
Černý, D and Simonoff, AL (2023) Statistical evaluation of character support reveals the instability of higher-level dinosaur phylogeny. Scientific Reports 13, 9273.CrossRefGoogle ScholarPubMed
Chiarenza, AA, Farnsworth, A, Mannion, PD, Lunt, DJ, Valdes, PJ, Morgan, JV and Allison, PA (2020) Asteroid impact, not volcanism, caused the end-Cretaceous dinosaur extinction. Proceedings of the National Academy of Sciences 117, 1708417093.CrossRefGoogle Scholar
Chiarenza, AA, Mannion, PD, Lunt, DJ, Farnsworth, A, Jones, LA, Kelland, S-J and Allison, PA (2019) Ecological niche modelling does not support climatically-driven dinosaur diversity decline before the Cretaceous/Paleogene mass extinction. Nature Communications 10, 114.CrossRefGoogle Scholar
Close, RA, Benson, RBJ, Alroy, J, Behrensmeyer, AK, Benito, J, Carrano, MT, Clearly, TJ, Dunne, EM, Mannion, PD, Uhen, MD and Butler, RJ (2019) Diversity dynamics of Phanerozoic terrestrial tetrapods at the local-community scale. Nature Ecology & Evolution 3, 590597.CrossRefGoogle ScholarPubMed
Close, RA, Benson, RBJ, Upchurch, P and Butler, RJ (2017) Controlling for the species-area effect supports constrained long-term Mesozoic terrestrial vertebrate diversification. Nature Communications 8, 15381.CrossRefGoogle ScholarPubMed
Close, RA, Evers, SW, Alroy, J and Butler, RJ (2018) How should we estimate diversity in the fossil record? Testing richness estimators using sampling-standardised discovery curves. Methods in Ecology and Evolution 9, 13861400.CrossRefGoogle Scholar
Cohen, KM, Finney, SC, Gibbard, PL and Fan, JX (2013) The ICS international chronostratigraphic chart. Episodes 36, 199204.CrossRefGoogle Scholar
Condamine, FL, Guinot, G, Benton, MJ and Currie, PJ (2021) Dinosaur biodiversity declined well before the asteroid impact, influenced by ecological and environmental pressures. Nature Communications 12, 3833.CrossRefGoogle ScholarPubMed
Dean, CD, Chiarenza, AA and Maidment, SCR (2020) Formation binning: A new method for increased temporal resolution in regional studies, applied to the late Cretaceous dinosaur fossil record of North America. Palaeontology 63, 881901.CrossRefGoogle Scholar
Dunhill, AM, Hannisdal, B, Brocklehurst, N and Benton, MJ (2018) On formation‐based sampling proxies and why they should not be used to correct the fossil record. Palaeontology 61, 119132.CrossRefGoogle Scholar
Fastovsky, DE, Huang, YF, Hsu, J, Martin-McNaughton, J, Sheehan, PM and Weishampel, DB (2004). Shape of Mesozoic dinosaur richness. Geology 32, 877880.CrossRefGoogle Scholar
García-Girón, J, Chiarenza, AA, Alahuhta, J, DeMar, J, Heino, J, Mannion, PD, Williamson, TE, Wilson Mantilla, GP and Brusatte, SL (2022) Shifts in food webs and niche stability shaped survivorship and extinction at the end-Cretaceous. Science Advances 8, eadd5040.CrossRefGoogle ScholarPubMed
Gavryushkina, A, Welch, D, Stadler, T and Drummond, AJ (2014) Bayesian inference of sampled ancestor trees for epidemiology and fossil calibration. PLoS Computational Biology 10, e1003919.CrossRefGoogle ScholarPubMed
Grenfell, BT, Pybus, OG, Gog, JR, Wood, JLN, Daly, JM, Mumford, JA and Holmes, EC (2004) Unifying the epidemiological and evolutionary dynamics of pathogens. Science 303, 327332.CrossRefGoogle ScholarPubMed
Griffiths, RC and Tavaré, S (1994) Sampling theory for neutral alleles in a varying environment. Philosophical Transactions of the Royal Society B 344, 403410.Google Scholar
Hadfield, JD (2010) MCMC methods for multi-response generalized linear mixed models: The MCMCglmm R package. Journal of Statistical Software 33, 122.CrossRefGoogle Scholar
Han, F, Wang, Q, Wang, H-P, Zhu, X-F, Zhou, X-Y, Wang, Z-X, Fang, K-Y, Stidham, TA, Wang, W, Wang, X-L, Li, X-Q, Qin, H-F, Fan, L-G, Wen, C, Luo, J-H, Pan, Y-X and Deng, C-L (2022) Low dinosaur biodiversity in central China 2 million years prior to the end-Cretaceous mass extinction. Proceedings of the National Academy of Sciences USA 119, e2211234119.CrossRefGoogle Scholar
Heath, TA, Huelsenbeck, JP and Stadler, T (2014) The fossilized birth-death process for coherent calibration of divergence-time estimates. Proceedings of the National Academy of Sciences USA 111, E2957E2966.CrossRefGoogle ScholarPubMed
Henao Diaz, LF, Harmon, LJ, Sugawara, MTC, Miller, ET and Pennell, MW (2019) Macroevolutionary diversification rates show time dependency. Proceedings of the National Academy of Sciences 116, 74037408.CrossRefGoogle ScholarPubMed
Hull, PM, Bornemann, A, Penman, DE, Henehan, MJ, Norris, RD, Wilson, PA, Blum, P, Alegret, L, Batenburg, SJ, Bown, PR, Bralower, TJ, Cournede, C, Deutsch, A, Donner, B, Friedrich, O, Jehle, S, Kim, H, Kroon, D, Lippert, PC, Loroch, D, Moebius, I, Moriya, K, Peppe, DJ, Ravizza, GE, Röhl, U, Schueth, JD, Sepúlveda, J, Sexton, PF, Sibert, EC, Śliwińska, KK, Summons, RE, Thomas, E, Westerhold, T, Whiteside, JH, Yamaguchi, T and Zachos, JC (2020) On impact and volcanism across the Cretaceous-Paleogene boundary. Science 367, 266272.CrossRefGoogle ScholarPubMed
Hurlbert, SH and Archibald, JD (1995) No statistical support for sudden (or gradual) extinction of dinosaurs. Geology 23, 881884.2.3.CO;2>CrossRefGoogle Scholar
Jones, LA, Gearty, W, Allen, BJ, Eichenseer, K, Dean, CD, Galván, S, Kouvari, M, Godoy, PL, Nicholl, C, Buffan, L, Flannery-Sutherland, JT, Dillon, EM and Chiarenza, AA (2023) palaeoverse: A community-driven R package to support palaeobiological analysis. Methods in Ecology and Evolution 14, 22052215.CrossRefGoogle Scholar
Karcher, MD, Palacios, JA, Bedford, T, Suchard, MA and Minin, VN (2016) Quantifying and mitigating the effect of preferential sampling on phylodynamic inference. PLoS Computational Biology 12, e1004789.CrossRefGoogle ScholarPubMed
Kingman, JFC (1982) The coalescent. Stochastic Processes and their Applications 13, 235248.CrossRefGoogle Scholar
Kühnert, D, Stadler, T, Vaughan, TG and Drummond, AJ (2016) Phylodynamics with migration: A computational framework to quantify population structure from genomic data. Molecular Biology and Evolution 33, 21022116.CrossRefGoogle ScholarPubMed
Langer, MC and Godoy, PL (2022) So volcanoes created the dinosaurs? A quantitative characterization of the early evolution of terrestrial Pan-Aves. Frontiers in Earth Science 10, 899562.CrossRefGoogle Scholar
Le Loeuff, J (2012) Paleobiogeography and biodiversity of late Maastrichtian dinosaurs: How many dinosaur species went extinct at the Cretaceous-Tertiary boundary? Bulletin de La Société Géologique de France 183, 547559.CrossRefGoogle Scholar
Legried, B and Terhorst, J (2022) A class of identifiable phylogenetic birth–death models. Proceedings of the National Academy of Sciences USA 119, e2119513119.CrossRefGoogle ScholarPubMed
Lloyd, GT (2011) A refined modelling approach to assess the influence of sampling on palaeobiodiversity curves: New support for declining Cretaceous dinosaur richness. Biology Letters 8, 123126.CrossRefGoogle ScholarPubMed
Lloyd, GT, Bapst, DW, Friedman, M and Davis, KE (2017) Probabilistic divergence time estimation without branch lengths: Dating the origins of dinosaurs, avian flight and crown birds. Biology Letters 12, 20160609.CrossRefGoogle Scholar
Lloyd, GT, Davis, KE, Pisani, D, Tarver, JE, Ruta, M, Sakamoto, M, Hone, DWE, Jennings, R and Benton, MJ (2008) Dinosaurs and the Cretaceous terrestrial revolution. Proceedings of the Royal Society B 275, 24832490.CrossRefGoogle ScholarPubMed
Louca, S and Pennell, MW (2020) Extant timetrees are consistent with a myriad of diversification histories. Nature 580, 502505.CrossRefGoogle ScholarPubMed
Mannion, PD, Upchurch, P, Carrano, MT and Barrett, PM (2011) Testing the effect of the rock record on diversity: A multidisciplinary approach to elucidating the generic richness of sauropodomorph dinosaurs through time. Biological Reviews 86, 157181.CrossRefGoogle ScholarPubMed
Marshall, CR, Finnegan, S, Clites, EC, Holroyd, PA, Bonuso, N, Cortez, NC, Davis, E, Dietl, GP, Druckenmiller, PS, Eng, RC, Garcia, C, Estes-Smargiassi, K, Hendy, A, Hollis, KA, Little, H, Nesbitt, EA, Roopnarine, P, Skibinski, L, Vendetti, J and White, LD (2018) Quantifying the dark data in museum fossil collections as palaeontology undergoes a second digital revolution. Biology Letters 14, 20180431.CrossRefGoogle ScholarPubMed
Mitchell, JS, Roopnarine, PD and Angielczyk, KD (2012) Late Cretaceous restructuring of terrestrial communities facilitated the end-Cretaceous mass extinction in North America. Proceedings of the National Academy of Sciences USA 109, 1885718861.CrossRefGoogle ScholarPubMed
O’Reilly, JE and Donoghue, PCJ (2020) Effect of fossil sampling on the estimation of divergence times with the fossilized birth-death process. Systematic Biology 69, 124138.CrossRefGoogle ScholarPubMed
Paradis, E and Schliep, K (2019) ape 5.0: An environment for modern phylogenetics and evolutionary analyses in R. Bioinformatics 35, 526528.CrossRefGoogle ScholarPubMed
R Core Team (2022) R: A Language and Environment for Statistical Computing. Vienna, Austria: R Foundation for Statistical Computing. Available at https://www.R-project.org/.Google Scholar
Raja, NB, Dunne, EM, Matiwane, A, Khan, TM, Nätscher, PS, Ghilardi, AM and Chattopadhyay, D (2022) Colonial history and global economics distort our understanding of deep-time biodiversity. Nature Ecology and Evolution 6, 145154.CrossRefGoogle ScholarPubMed
Rambaut, A and Drummond, AJ (2021) TreeAnnotator (version 2.6.6). Available at https://beast.community/treeannotator.Google Scholar
Raup, DM (1976) Species diversity in the Phanerozoic: A tabulation. Paleobiology 2, 279288.CrossRefGoogle Scholar
Russell, DA (1995) China and the lost worlds of the dinosaurian era. Historical Biology 10, 312.CrossRefGoogle Scholar
Sakamoto, M, Benton, MJ and Venditti, C (2016) Dinosaurs in decline tens of millions of years before their final extinction. Proceedings of the National Academy of Sciences 113, 50365040.CrossRefGoogle ScholarPubMed
Sakamoto, M, Benton, MJ and Venditti, C (2021) Strong support for a heterogeneous speciation decline model in Dinosauria: A response to claims made by Bonsor et al. (2020). Royal Society Open Science 8, 202143.CrossRefGoogle ScholarPubMed
Sakamoto, M, Venditti, C and Benton, MJ (2017) ‘Residual diversity estimates’ do not correct for sampling bias in palaeodiversity data. Methods in Ecology and Evolution 8, 453459.CrossRefGoogle Scholar
Schulte, P, Alegret, L, Arenillas, I, Arz, JA, Barton, PJ, Bown, PR, Bralower, TJ, Christeson, GL, Claeys, P, Cockell, CS, Collins, GS, Deutsch, A, Goldin, TJ, Goto, K, Grajales-Nishimura, JM, Grieve, RA, Gulick, SPS, Johnson, KR, Kiessling, W, Koeberl, C, Kring, DA, MacLeod, KG, Matsui, T, Melosh, J, Montanari, A, Morgan, JV, Neal, CR, Nichols, DJ, Norris, RD, Pierazzo, E, Ravizza, G, Rebolledo-Vieyra, M, Reimold, WU, Robin, E, Salge, T, Speijer, RP, Sweet, AR, Urrutia-Fucugauchi, J, Vajda, V, Whale, MT and Willumsen, PS (2010) The Chicxulub asteroid impact and mass extinction at the Cretaceous-Paleogene boundary. Science 327, 12141218.CrossRefGoogle ScholarPubMed
Signor, PW and Lipps, JH (1982) Sampling bias, gradual extinction patterns and catastrophes in the fossil record. Geological Society of America Special Paper 190, 291296.CrossRefGoogle Scholar
Silvestro, D, Salamin, N and Schnitzler, J (2014) PyRate: A new program to estimate speciation and extinction rates from incomplete fossil data. Methods in Ecology and Evolution 5, 11261131.CrossRefGoogle Scholar
Sloan, RE, Rigby, JK Jr., Van Valen, LM and Gabriel, D (1986) Gradual dinosaur extinction and simultaneous ungulate radiation in the Hell Creek formation. Science 232, 629633.CrossRefGoogle ScholarPubMed
Smiley, TM (2018) Detecting diversification rates in relation to preservation and tectonic history from simulated fossil records. Paleobiology 44, 124.CrossRefGoogle Scholar
Stadler, T (2011) Mammalian phylogeny reveals recent diversification rate shifts. Proceedings of the National Academy of Sciences USA 108, 61876192.CrossRefGoogle ScholarPubMed
Stadler, T, Gavryushkina, A, Warnock, RCM, Drummond, AJ and Heath, TA (2018) The fossilized birth-death model for the analysis of stratigraphic range data under different speciation modes. Journal of Theoretical Biology 447, 4155.CrossRefGoogle ScholarPubMed
Stadler, T, Kühnert, D, Bonhoeffer, S and Drummond, AJ (2013) Birth-death skyline plot reveals temporal changes of epidemic spread in HIV and hepatitis C virus (HCV). Proceedings of the National Academy of Sciences USA 110, 228233.CrossRefGoogle ScholarPubMed
Starrfelt, J and Liow, LH (2016) How many dinosaur species were there? Fossil bias and true richness estimated using a Poisson sampling model. Philosophical Transactions of the Royal Society B 371, 20150219.CrossRefGoogle ScholarPubMed
Truman, K, Vaughan, TG, Gavryushkin, A and Gavryushkina, A (2024) The fossilised birth-death model is identifiable. bioRxiv.CrossRefGoogle Scholar
Uhen, MD, Allen, B, Behboudi, N, Clapham, ME, Dunne, E, Hendy, A, Holroyd, PA, Hopkins, M, Mannion, P, Novack-Gottshall, P, Pimiento, C and Wagner, P (2023) Paleobiology Database User Guide Version 1.0. PaleoBios 40, 156.CrossRefGoogle Scholar
Upchurch, P, Mannion, PD, Benson, RBJ, Butler, RJ and Carrano, MT (2011) Geological and anthropogenic controls on the sampling of the terrestrial fossil record: A case study from the Dinosauria. Geological Society London Special Publications 358, 209240.CrossRefGoogle Scholar
Walker, FM, Dunhill, AM and Benton, MJ (2020) Variable preservation potential and richness in the fossil record of vertebrates. Palaeontology 63, 313329.CrossRefGoogle Scholar
Wang, SC and Dodson, P (2006) Estimating the diversity of dinosaurs. Proceedings of the National Academy of Sciences USA 103, 1360113605.CrossRefGoogle ScholarPubMed
Warnock, RCM, Heath, TA and Stadler, T (2020) Assessing the impact of incomplete species sampling on estimates of speciation and extinction rates. Paleobiology 46, 137157.CrossRefGoogle Scholar
Figure 0

Table 1. Priors for the Bayesian phylodynamic analyses

Figure 1

Figure 1. Diversification rates estimated using the piecewise-exponential coalescent model. Time moves forwards from left to right along the x-axis, with the K-Pg boundary at the end of the Coniacian–Maastrichtian bin. Estimates are shown for each of four phylogenies, ordered from smallest to largest. Points show the median values, and error bars indicate 95% highest posterior density. Dinosaur silhouettes for Ornithischia (top right), Sauropodomorpha (bottom left) and Theropoda (bottom right) are from Phylopic.

Figure 2

Figure 2. Diversification rates estimated using the piecewise-constant fossilised birth-death skyline model. Time moves forwards from left to right along the x-axis, with the K-Pg boundary at the end of the Coniacian–Maastrichtian bin. Estimates are shown for each of four phylogenies, ordered from smallest to largest. Points show the median values, and error bars indicate 95% highest posterior density. Dinosaur silhouettes for Ornithischia (top right), Sauropodomorpha (bottom left) and Theropoda (bottom right) are from Phylopic.

Figure 3

Figure 3. Sampling rates estimated using the piecewise-constant fossilised birth-death model. Time moves forwards from left to right along the x-axis, with the K-Pg boundary at the end of the Coniacian–Maastrichtian bin. Estimates are shown for each of four phylogenies, ordered from smallest to largest. Points show the median values, and error bars indicate 95% highest posterior density. Dinosaur silhouettes for Ornithischia (top right), Sauropodomorpha (bottom left) and Theropoda (bottom right) are from Phylopic.

Supplementary material: File

Allen et al. supplementary material

Allen et al. supplementary material
Download Allen et al. supplementary material(File)
File 499.7 KB

Author comment: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R0/PR1

Comments

No accompanying comment.

Recommendation: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R0/PR2

Comments

EXT-22-0017 „Mechanistic phylodynamic models do not provide conclusive evidence that dinosaurs were in decline before their final extinction”

While reviewer #1 was happy about your submission and suggested only minor revisions (mostly regarding colloquial language), reviewer #2 was more critical along diverse routes. Below, I provide my take on reviewer’s #2 critique:

Your phylogenetic dataset may not be up to date (a view shared with reviewer #1) and simple tree-averaging might not be justified. I understand that it is beyond the scope of this paper to produce a new supertree of dinosaurs. However, it would indeed be good to use the state-of-the-art tree and I know too little of dinosaurs to judge whether a more recent tree is available. Even if it is not, it may indeed be better to give more weight to a more recent tree in case of conflicting results. The suggestion to see how results depend on phylogenies vs. phylodynamic models is a good one and should be pursued by the authors.

Regarding novelty, I share the reviewers concern that the basic conclusion have been reached many time before. Quantifying the decline or absence thereof might be a worthwhile endeavor.

Decision: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R0/PR3

Comments

No accompanying comment.

Author comment: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R1/PR4

Comments

No accompanying comment.

Recommendation: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R1/PR5

Comments

This is a well-written and thought-provoking manuscript arguing that the available data (and sampling models) are simply not sufficient to judge if non-avian dinosaurs were in decline prior to their extinction at the end of the Cretaceous period.

The reviews suggest (very) minor revisions and I second their evaluation. Actually, while a short comment on Louca & Pennell (2020, Nature 580:502-505) could be useful, I ask the authors to refrain from a lengthy discussion about the use of timetrees in diversification dynamics (suggested by reviewer #1). The Signor-Lipps effect on the branch ends, however, is a valid point that should be tackled. Perhaps this is already considered in the model.

The only major comment of reviewer #2 regards the temporal binning of the model (change times). I second that this appears a bit ad hoc and requires justification.

l. 91-92: “infer what we do not know from what we do.” I don’t understand the “from what we do” subphrase

Figs. 1-3: The dark grey vertical stripes should be lighter. The black (Lloyd1) bars are almost not visible. Yes transforming the y axis in Fig. 3 might be useful, although a square-root transformation might work better than a log transformation as you seem to have several 0 values.

Decision: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R1/PR6

Comments

No accompanying comment.

Author comment: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R2/PR7

Comments

No accompanying comment.

Recommendation: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R2/PR8

Comments

The authors have successfully dealt with all pending issues and the manuscript is now in perfect shape to be accepted for our journal. Thank you for this important study.

Decision: Mechanistic phylodynamic models do not provide conclusive evidence that non-avian dinosaurs were in decline before their final extinction — R2/PR9

Comments

No accompanying comment.