Hostname: page-component-76fb5796d-qxdb6 Total loading time: 0 Render date: 2024-04-27T10:29:10.045Z Has data issue: false hasContentIssue false

The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future

Published online by Cambridge University Press:  07 June 2023

Peter Wilf*
Affiliation:
Department of Geosciences and Earth and Environmental Systems Institute, Pennsylvania State University, University Park, PA, USA
Mónica R. Carvalho
Affiliation:
Museum of Paleontology and Department of Earth and Environmental Sciences, University of Michigan, Ann Arbor, MI, USA
Elena Stiles
Affiliation:
Department of Biology, University of Washington, Seattle, WA, USA
*
Corresponding author: Peter Wilf; Email: pwilf@psu.edu
Rights & Permissions [Opens in a new window]

Abstract

The Cretaceous–Paleogene (K–Pg) mass extinction was geologically instantaneous, causing the most drastic extinction rates in Earth’s History. The rapid species losses and environmental destruction from the Chicxulub impact at 66.02 Ma made the K–Pg the most comparable past event to today’s projected “sixth” mass extinction. The extinction famously eliminated major clades of animals and plankton. However, for land plants, losses primarily occurred among species observed in regional studies but left no global trace at the family or major-clade level, leading to questions about whether there was a significant K–Pg plant extinction. We review emerging paleobotanical data from the Americas and argue that the evidence strongly favors profound (generally >50%), geographically heterogeneous species losses and recovery consistent with mass extinction. The heterogeneity appears to reflect several factors, including distance from the impact site and marine and latitudinal buffering of the impact winter. The ensuing transformations have affected all land life, including true angiosperm dominance in the world’s forests, the birth of the hyperdiverse Neotropical rainforest biome, and evolutionary radiations leading to many crown angiosperm clades. Although the worst outcomes are still preventable, the sixth mass extinction could mirror the K–Pg event by eliminating comparable numbers of plant species in a geologic instant, impoverishing and eventually transforming terrestrial ecosystems while having little effect on global plant-family diversity.

Type
Review
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), 2023. Published by Cambridge University Press

Impact statement

The impact of an asteroid the size of San Francisco with Yucatán, Mexico at the end of the Cretaceous period, 66 million years ago, set a devastating series of events in motion. The massive species losses, popularly known as the “dinosaur extinction,” eliminated an estimated 75% of land and sea species. The end-Cretaceous is highly relevant to modern-day projected extinctions as the only event in Earth’s history that destroyed environments and wiped out life worldwide in a geologic instant. Land plants are the foundation of life on land, but suitable fossil-plant collections from the critical time interval are rare and geographically biased, leading to debate about whether there was any significant global plant extinction. We review new records of plant fossils from North and South America, which indicate significant (>50%) plant-species losses in each area, consistent with mass extinction. The different regions demonstrate heterogeneous responses, which are plausibly related to variations in distance to Mexico and the severity of the impact winter caused by atmospheric particles. Freezing the tropics, the most biodiverse region, may have generated devastating species losses. After the disaster came transformations of terrestrial ecosystems that define life on land today, including dramatic crown-group radiations, the rise to true dominance of flowering plants, and the birth of hyperdiverse tropical rainforests. Although it is not too late to avert the worst outcomes, projected losses of plant species in the near future are similar to estimates from the end-Cretaceous extinction. Ecological transformations will eventually follow, in all likelihood, too late to benefit humans.

Introduction

Critical time intervals provide insights into the projected “sixth mass extinction” from anthropogenic disturbances (Wake and Vredenburg, Reference Wake and Vredenburg2008; Barnosky et al., Reference Barnosky, Matzke, Tomiya, Wogan, Swartz, Quental, Marshall, McGuire, Lindsey, Maguire, Mersey and Ferrer2011; Wing and Currano, Reference Wing and Currano2013). However, the 66.02 Ma Cretaceous–Paleogene (K–Pg) mass extinction (here, KPgE) or “dinosaur extinction” stands above all known biotic crises for its suddenness. The KPgE destroyed approximately 75% of marine species and similar numbers of land animals in a geologic instant (see Jones et al., Reference Jud and Gandolfo2023 for a recent summary). Of the “Big Five” mass extinctions (Raup and Sepkoski, Reference Raup and Sepkoski1982; Marshall, Reference Marshall2023) and other critical intervals (Kiessling et al., Reference Kiessling, Smith and Raja2023), only the KPgE is directly comparable with the modern day for its speed of killing and environmental destruction (Barnosky et al., Reference Barnosky, Matzke, Tomiya, Wogan, Swartz, Quental, Marshall, McGuire, Lindsey, Maguire, Mersey and Ferrer2011).

Since Alvarez et al. (Reference Alvarez, Alvarez, Asaro and Michel1980) proposed their impact-extinction hypothesis for the KPgE, the supporting evidence has grown overwhelmingly (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, Whalen and Willumsen2010; 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). The unique cause of the KPgE was the Chicxulub bolide impact in Yucatán, Mexico at 66.02 Ma. The collision generated a global event horizon, an era boundary preserved at hundreds of sites from land to the deep ocean (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, Whalen and Willumsen2010; Morgan et al., Reference Morgan, Bralower, Brugger and Wünnemann2022). Owing to recent drilling, the 200-km-wide Chicxulub crater’s structure is known in outstanding detail, providing a highly resolved history of the grim “first day of the Cenozoic” (Gulick et al., Reference Gulick, Bralower, Ormö, Hall, Grice, Schaefer, Lyons, Freeman, Morgan, Artemieva, Kaskes, de Graaff, Whalen, Collins, Tikoo, Verhagen, Christeson, Claeys, Coolen, Goderis, Goto, Grieve, McCall, Osinski, Rae, Riller, Smit, Vajda and Wittmann2019; Morgan et al., Reference Morgan, Bralower, Brugger and Wünnemann2022). The crater’s morphology is comparable to the largest observed in the solar system (Morgan et al., Reference Morgan, Bralower, Brugger and Wünnemann2022). The estimated impact energy is 1023 joules, and returning tsunamis brought charcoal from distant forest fires back into the crater (Gulick et al., Reference Gulick, Bralower, Ormö, Hall, Grice, Schaefer, Lyons, Freeman, Morgan, Artemieva, Kaskes, de Graaff, Whalen, Collins, Tikoo, Verhagen, Christeson, Claeys, Coolen, Goderis, Goto, Grieve, McCall, Osinski, Rae, Riller, Smit, Vajda and Wittmann2019). The sky went dark for months to years from dust, iron-oxide nanoparticles, aerosols, and 1014–15g of black carbon from the target rock entering the upper atmosphere (Vajda et al., Reference Vajda, Ocampo, Ferrow and Koch2015; Lyons et al., Reference Lyons, Karp, Bralower, Grice, Schaefer, Gulick, Morgan and Freeman2020), thus freezing the surface, stopping photosynthesis, and causing the bulk of the mass extinctions.

Land plants are the foundational macroorganisms of terrestrial ecosystems, and their fate at the KPgE has drawn attention since well before 1980 (Dorf, Reference Dorf1940; see Nichols and Johnson, Reference Nichols and Johnson2008). However, plants show very different extinction patterns from animals, partly because of their dramatically different life histories (Traverse, Reference Traverse1988; Wing, Reference Wing and Taylor2004; Cascales-Miñana and Cleal, Reference Cascales-Miñana and Cleal2014). The plant-fossil record since the Late Cretaceous is dominated by angiosperm leaves, which are well suited for species-level quantitative analyses in regional studies. However, most are not formally or accurately described and cannot be easily assigned to higher taxa, restricting spatial comparability (Wilf, Reference Wilf2008). The result is a sharp contrast between high-resolution regional stratigraphic studies, which usually show significant plant-species extinction and ecological shifts, and global syntheses or phylogenetic analyses, which show no losses of plant families or major clades. This situation has led to a debate regarding whether there was any significant K–Pg extinction for plants (Cascales-Miñana and Cleal, Reference Cascales-Miñana and Cleal2014; Nic Lughadha et al., Reference Nic Lughadha, Bachman, Leão, Forest, Halley, Moat, Acedo, Bacon, Brewer, Gâteblé, Gonçalves, Govaerts, Hollingsworth, Krisai‐Greilhuber, de Lirio, Moore, Negrão, Onana, Rajaovelona, Razanajatovo, Reich, Richards, Rivers, Cooper, Iganci, Lewis, Smidt, Antonelli, Mueller and Walker2020; Thompson and Ramírez-Barahona, Reference Thompson and Ramírez-Barahona2023). Several previous reviews have covered the general topics of K–Pg plant extinctions, the vital contributions of palynology, and the rich history of investigations (Wing, Reference Wing and Taylor2004; McElwain and Punyasena, Reference McElwain and Punyasena2007; Nichols and Johnson, Reference Nichols and Johnson2008; Vajda and Bercovici, Reference Vajda and Bercovici2014).

Here, we focus on the improving macrofossil (and associated palynological) records of the KPgE from the Americas and the increasing understanding of heterogeneity in the event’s severity, ecosystem effects, and legacies. Despite advances in many regions, only the Western Hemisphere has extensive, stratigraphically well-constrained collections of latest Cretaceous and early Paleocene plant macrofossils (e.g., Nichols and Johnson, Reference Nichols and Johnson2008; Vajda and Bercovici, Reference Vajda and Bercovici2014). We find significant evidence for species losses of land plants that fulfill any reasonable definition of mass extinction and provide instructive analogs for the near future.

Heterogeneity

General regional differences in the K–Pg plant extinction and recovery are well known (Wolfe, Reference Wolfe1987; Nichols and Johnson, Reference Nichols and Johnson2008; Vajda and Bercovici, Reference Vajda and Bercovici2014). However, a pronounced North American sampling bias exists for nearly all significant collections of Cretaceous (K floras) and Paleocene floras from well-dated stratigraphic sections with well-defined K–Pg event layers (Nichols and Johnson, Reference Nichols and Johnson2008). The Williston Basin in the northern Great Plains, USA, paleolatitude ca. 50°N, remains by far the best-studied area (Johnson et al., Reference Johnson, Nichols, Attrep and Orth1989; Johnson, Reference Johnson2002).

Several factors are likely to promote heterogeneity. There is a global pattern of increasing K–Pg event layer thickness and sediment disturbance with proximity to the Chicxulub crater (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, Whalen and Willumsen2010). Thus, the first effect usually considered is the distance from ground zero and the proximal effects of shockwaves, tsunamis, and large ejecta. A second factor highly relevant to plants is maritime and latitudinal buffering of the impact winter (Figure 1; Bardeen et al., Reference Bardeen, Garcia, Toon and Conley2017; Brugger et al., Reference Brugger, Feulner and Petri2017; Morgan et al., Reference Morgan, Bralower, Brugger and Wünnemann2022). Both distance and buffering gradients predict large extinctions in western North America and the Neotropics, especially if the tropics froze, and less severe extinctions in the maritime areas of temperate Gondwana, where there is a growing list of survivor taxa (e.g., McLoughlin et al., Reference McLoughlin, Carpenter and Pott2011; summarized in Wilf et al., Reference Wilf, Cúneo, Escapa, Pol and Woodburne2013). Biotic variables include the diversity and traits (Berry, Reference Berry2020; Butrim et al., Reference Butrim, Royer, Miller, Dechesne, Neu-Yagle, Lyson, Johnson and Barclay2022) of pre-impact and recovery floras.

Figure 1. Modeled surface air temperatures (A) before and (B) during the coldest impact-winter year, 3 years after the Chicxulub impact; pickaxe icons indicate the principal fossil floras discussed in the text from (north to south) the northern Great Plains, southern Rockies, Colombia, and southern Argentina. Drafted by J. Brugger, using model output from Brugger et al. (Reference Brugger, Feulner, Hofmann and Petri2021), with permission.

North America

In the Williston Basin near Marmarth, North Dakota, more than 160 plant-macrofossil localities spanning ca. 67.4–65.2 Ma were placed in precise stratigraphic positions measured to the well-preserved K–Pg event bed (Johnson et al., Reference Johnson, Nichols, Attrep and Orth1989; Johnson, Reference Johnson2002). The total number of paleobotanical specimens exceeds 22,000, with a significant representation of K and Paleocene floras. A conservative, robust estimate of macrofloral species extinction from these collections is 57%, based only on the species present in the uppermost 5 m (approximately 70 Kyr) of Cretaceous strata (Wilf and Johnson, Reference Wilf and Johnson2004; Figure 2A). No other location has comparable sampling density and temporal precision, which must be considered when comparing extinction percentages (Figure 2B). Similar patterns of severe plant-species extinction, low early Paleocene diversity (“disaster floras”), and ca. 10 million years of recovery are known from significant but less complete datasets throughout the northern Great Plains (Wing et al., Reference Wing, Alroy and Hickey1995; Hotton, Reference Hotton2002; Nichols and Johnson, Reference Nichols and Johnson2008; Wilson et al., Reference Wilson, Wilson Mantilla and Strӧmberg2021). The regional Paleocene floras are well understood systematically and include many lineages that remain abundant in living north-temperate floras, such as taxodioid Cupressaceae, Cornales, Fagales, and Platanaceae (Brown, Reference Brown1962; Hickey, Reference Hickey1977; Manchester, Reference Manchester2014).

Figure 2. Plant extinction in the Williston Basin, North Dakota. (A) Ranges in the composite section of 141 macrofossil species occurring in more than one stratigraphic bin, with 99% confidence intervals on last appearances, relative to the K/Pg event horizon recognized from multiple indicators. The conservative estimate of 57% species extinction is based on this analysis, using only the last appearances in the uppermost 5 m of Cretaceous strata (Wilf and Johnson, Reference Wilf and Johnson2004). The observed extinction increases if the window is shifted down, as demonstrated in (B), showing simulated extinctions and species richness from the same dataset (singletons included) when based on discrete 10-m bin windows (Stiles et al., Reference Stiles, Wilf, Iglesias, Gandolfo and Cúneo2020). Modified from (A) Wilf and Johnson (Reference Wilf and Johnson2004) and (B) Stiles et al. (Reference Stiles, Wilf, Iglesias, Gandolfo and Cúneo2020).

The Southern Rockies present an emerging record of early Paleocene sites, several of which have much higher plant diversity than the northern Great Plains; however, few latest Cretaceous floras are known for comparison. The region is closer than the Williston Basin to Chicxulub, refuting distance from the crater as an explanation for the elevated Paleocene diversity and pointing to latitudinal effects combined with lower continentality (Figure 1B). Some sites in Colorado and New Mexico are said to preserve “tropical” forests (Johnson and Ellis, Reference Johnson and Ellis2002; Flynn and Peppe, Reference Flynn and Peppe2019), despite their location in temperate latitudes. This use of “tropical” to refer to warm past climates at middle latitudes is common in paleontology; however, the term is best applied only to the latitudinal tropics (<23.5°), where insolation and its effects on plant life are far more significant (e.g., Jaramillo and Cárdenas, Reference Jaramillo and Cárdenas2013).

Flynn and Peppe (Reference Flynn and Peppe2019) reported diverse early Paleocene (<350 Kyr after the K–Pg) floras from the San Juan Basin in New Mexico; there is no event horizon preserved, nor are K floras available for comparison. These assemblages represent frost-free conditions but lack characteristic tropical plant taxa (e.g., Carvalho et al., Reference Carvalho, Jaramillo, de la Parra, Caballero-Rodríguez, Herrera, Wing, Turner, D’Apolito, Romero-Báez, Narváez, Martínez, Gutierrez, Labandeira, Bayona, Rueda, Paez-Reyes, Cárdenas, Duque, Crowley, Santos and Silvestro2021). Instead, they appear to preserve a mixture of warm-temperate lineages similar to those found throughout the Paleocene north temperate zone, including species of Cupressaceae, Platanaceae, and Cornales. Further north, the Raton Basin of New Mexico and Colorado preserves a well-defined K–Pg event bed and was the primary location of early work on K–Pg plant extinctions following the publication of the Alvarez hypothesis (Tschudy et al., Reference Tschudy, Pillmore, Orth, Gilmore and Knight1984; Wolfe and Upchurch, Reference Wolfe and Upchurch1986). The area has received renewed attention, particularly concerning the basalmost Paleocene floras (Berry, Reference Berry2019, Reference Berry2023).

A rich macrofloral record comes from over 150 Late Cretaceous and Paleogene sites in the Denver Basin, Colorado (Johnson et al., Reference Johnson, Reynolds, Werth and Thomasson2003). The sites were stratigraphically positioned using a detailed basin age-depth model anchored to a well-defined K–Pg event horizon exposed at West Bijou Creek and found in the subsurface in the Kiowa Core (Barclay et al., Reference Barclay, Johnson, Betterton and Dilcher2003; Raynolds et al., Reference Raynolds, Johnson, Ellis, Dechesne and Miller2007; Clyde et al., Reference Clyde, Ramezani, Johnson, Bowring and Jones2016). A showcase geochronology study combined paleomagnetic stratigraphy and U–Pb dating of a series of volcanic ashes at West Bijou Creek, constraining the age of the K–Pg boundary to 66.021 ± 0.024 Ma (Clyde et al., Reference Clyde, Ramezani, Johnson, Bowring and Jones2016).

We note that the exact age control of the KPgE represents decades of advances in geochronologic methods and inter-laboratory cooperation (Gradstein et al., Reference Gradstein, Ogg, Schmitz and Ogg2012), in contrast to the continuing absence of geological benchmarks and standards that are needed to align molecular divergence dates with the international geologic time scale (Wilf and Escapa, Reference Wilf and Escapa2016). We remain skeptical of the substantial molecular-dating literature that assigns whole genome duplications (WGDs) in many plant lineages to an algorithmically-approximated “K”-“Pg” “boundary,” implying that WGDs were beneficial for survival or recovery (e.g., Vanneste et al., Reference Vanneste, Baele, Maere and Van De Peer2014; Lohaus and Van de Peer, Reference Lohaus and Van de Peer2016; Koenen et al., Reference Koenen, Ojeda, Bakker, Wieringa, Kidner, Hardy, Pennington, Herendeen, Bruneau and Hughes2021). This work also used few well-defined geological or phylogenetic criteria for fossil calibrations (contra Sauquet et al., Reference Sauquet, Ho, Gandolfo, Jordan, Wilf, Cantrill, Bayly, Bromham, Brown, Carpenter, Lee, Murphy, Sniderman and Udovicic2012; Ramírez-Barahona et al., Reference Ramírez-Barahona, Sauquet and Magallón2020) and followed the widespread practice of appropriating the language of geochronology despite the absence of analyzed rocks, thereby conflating different sources of evidence. At a minimum, molecular age estimates should use a distinctive labeling system (e.g., Figure 3B).

Figure 3. The post-KPgE rise of extant angiosperm families, seen in Colombian macrofossils and a molecular-phylogenetic reconstruction of stem- vs. crown-group age distributions. In (A), bars show proportions of species (with standard deviations) in a family in a compilation of 72 Neotropical vegetation plots; darker bars indicate the top-10 families that account for half of total stems. Red diamonds mark families present in a latest Cretaceous flora; blue circles show families recognized in two Paleocene floras. Modified from Carvalho et al. (Reference Carvalho, Jaramillo, de la Parra, Caballero-Rodríguez, Herrera, Wing, Turner, D’Apolito, Romero-Báez, Narváez, Martínez, Gutierrez, Labandeira, Bayona, Rueda, Paez-Reyes, Cárdenas, Duque, Crowley, Santos and Silvestro2021). (B) Frequency distributions and 95% kernel density curves of reconstructed mean stem (red) and crown (blue) angiosperm family ages, from an analysis calibrated using 238 well-constrained fossils. Subscript “m” indicates age estimates derived from molecular analyses rather than direct geological dating. Modified from Ramírez-Barahona et al. (Reference Ramírez-Barahona, Sauquet and Magallón2020), with permission.

Figure 4. Early Danian reproductive fossils from the Salamanca Formation, Chubut, Argentina. (A) Agathis immortalis (Araucariaceae) branch with leaves and axillary pollen cones (see Escapa et al., Reference Escapa, Iglesias, Wilf, Catalano, Caraballo-Ortiz and Rubén Cúneo2018). (B) Notiantha grandensis (Rhamnaceae) flowers side-by-side in transverse and longitudinal views (see Jud et al., Reference Jud, Gandolfo, Iglesias and Wilf2017). (C) Stephania psittaca (Menispermaceae) endocarp (see Jud et al., 2018b). Lacinipetalum spectabilum (Cunoniaceae) flower, detail of ovary surface covered in trichomes and entrapped pollen grains, glowing under epifluorescence (see Jud et al., 2018a). All the specimens are housed at Museo Paleontolόgico Egidio Feruglio, Trelew, Chubut. Credits: A, C, D: P. Wilf; B, N. Jud, used with permission.

Returning to the Denver Basin, many of its early Paleocene floras are similar in composition and have low species diversity, like those of the northern Plains. However, several intriguing exceptions exist (Johnson et al., Reference Johnson, Reynolds, Werth and Thomasson2003), including the Corral Bluffs and Castle Rock floras. The Corral Bluffs (Lyson et al., Reference Lyson, Miller, Bercovici, Weissenburger, Fuentes, Clyde, Hagadorn, Butrim, Johnson, Fleming, Barclay, Maccracken, Lloyd, Wilson, Krause and Chester2019) exposures contain a well-dated sequence of ca. 66.2–65.2 Ma floras that co-occur with well-preserved mammalian remains, allowing direct correlations between floral recovery stages and increases in mammalian body size. The event horizon is not preserved, but the ages of the fossils are well constrained from a U–Pb dated ash layer, paleomagnetic stratigraphy, and a characteristic earliest Paleocene fern-spore spike. The sampling of K floras at Corral Bluffs is limited but sufficient to show a loss of more than half of the angiosperm species, consistent with the Williston Basin. Taxonomic work is at an early stage, but an exciting occurrence is the oldest definitive fossil legume fruit (Fabaceae); most elements, including Juglandaceae, Platanaceae, and palms, appear to be consistent with typical early Paleocene floras.

The ca. 63.8 Ma Castle Rock flora (Johnson and Ellis, Reference Johnson and Ellis2002; Ellis et al., Reference Ellis, Johnson and Dunn2003; Erdei et al., Reference Erdei, Coiro, Miller, Johnson Kirk, Griffith and Murphy2019) was the first to break the mold of low-diversity, homogenous early Paleocene assemblages. The estimated richness of more than 100 species at Castle Rock is several times that of typical Paleocene floras from the western USA. Variable composition at the local scale (beta diversity) and prevalence of notably large leaves with drip tips indicate warm and humid conditions, leading to the flora’s initial description as a “tropical rainforest in Colorado” (Johnson and Ellis, Reference Johnson and Ellis2002). Large angiosperm leaves are easily fragmented and rarely preserved as fossils (e.g., Merkhofer et al., Reference Merkhofer, Wilf, Haas, Kooyman, Sack, Scoffoni and Cúneo2015), and they only occur in the modern world in humid, warm habitats (Wright et al., Reference Wright, Dong, Maire, Prentice, Westoby, Díaz, Gallagher, Jacobs, Kooyman, Law, Leishman, Niinemets, Reich, Sack, Villar, Wang and Wilf2017). Thus, Castle Rock represents a minimally transported assemblage that records small-scale compositional variation and a warm, at least seasonally wet environment (Ellis et al., Reference Ellis, Johnson and Dunn2003). These observations led to the hypothesis that vegetation adjacent to the Laramide Front Range received abundant orographic rainfall, which fostered high floral diversity (Johnson and Ellis, Reference Johnson and Ellis2002). An alternative idea would be that Castle Rock more faithfully preserves large leaves and diversity than other early Paleocene sites but does not represent fundamentally different source vegetation similar to a tropical rainforest. Taxonomic studies of the Castle Rock macroflora that would address this issue are limited (Erdei et al., Reference Erdei, Coiro, Miller, Johnson Kirk, Griffith and Murphy2019). The elements identified to date appear to represent widespread clades in temperate-zone Paleocene floras, such as Platanaceae, Malvaceae, Lauraceae, and rare cycads (Ellis et al., Reference Ellis, Johnson and Dunn2003; Erdei et al., Reference Erdei, Coiro, Miller, Johnson Kirk, Griffith and Murphy2019). The palynoflora from Castle Rock contains standard early Paleocene taxa for the region, and no distinctive tropical rainforest elements have been reported (Nichols and Fleming, Reference Nichols and Fleming2002).

Neotropics

The wet tropical forests of South America, Africa, and Southeast Asia hold most of the Earth’s biodiversity (Slik et al., Reference Slik, Arroyo-Rodríguez, Aiba, Alvarez-Loayza, Alves, Ashton, Balvanera, Bastian, Bellingham, van den Berg, Bernacci, da Conceição Bispo, Blanc, Böhning-Gaese, Boeckx, Bongers, Boyle, Bradford, Brearley, Breuer-Ndoundou Hockemba, Bunyavejchewin, Calderado Leal Matos, Castillo-Santiago, Catharino, Chai, Chen, Colwell, Chazdon, Clark, Clark, Clark, Culmsee, Damas, Dattaraja, Dauby, Davidar, DeWalt, Doucet, Duque, Durigan, Eichhorn, Eisenlohr, Eler, Ewango, Farwig, Feeley, Ferreira, Field, de Oliveira Filho, Fletcher, Forshed, Franco, Fredriksson, Gillespie, Gillet, Amarnath, Griffith, Grogan, Gunatilleke, Harris, Harrison, Hector, Homeier, Imai, Itoh, Jansen, Joly, de Jong, Kartawinata, Kearsley, Kelly, Kenfack, Kessler, Kitayama, Kooyman, Larney, Laumonier, Laurance, Laurance, Lawes, Amaral, Letcher, Lindsell, Lu, Mansor, Marjokorpi, Martin, Meilby, Melo, Metcalfe, Medjibe, Metzger, Millet, Mohandass, Montero, de Morisson Valeriano, Mugerwa, Nagamasu, Nilus, Ochoa-Gaona, Onrizal Page, Parolin, Parren, Parthasarathy, Paudel, Permana, Piedade, Pitman, Poorter, Poulsen, Poulsen, Powers, Prasad, Puyravaud, Razafimahaimodison, Reitsma, dos Santos, Roberto Spironello, Romero-Saltos, Rovero, Rozak, Ruokolainen, Rutishauser, Saiter, Saner, Santos, Santos, Sarker, Satdichanh, Schmitt, Schöngart, Schulze, Suganuma, Sheil, da Silva Pinheiro, Sist, Stevart, Sukumar, Sun, Sunderland, Suresh, Suzuki, Tabarelli, Tang, Targhetta, Theilade, Thomas, Tchouto, Hurtado, Valencia, van Valkenburg, van do, Vasquez, Verbeeck, Adekunle, Vieira, Webb, Whitfeld, Wich, Williams, Wittmann, Wöll, Yang, Adou Yao, Yap, Yoneda, Zahawi, Zakaria, Zang, de Assis, Garcia Luize and Venticinque2015; Pillay et al., Reference Pillay, Venter, Aragon‐Osejo, González‐Del‐Pliego, Hansen, Watson and Venter2022); however, the effects of the KPgE in the tropics have been largely unknown. The situation changed with investigations of three outstanding macrofloras in Colombia, from the late Maastrichtian (ca. 68 Ma) Guaduas Formation and the middle-late Paleocene (ca. 60–58 Ma) Cerrejón and Bogotá formations (e.g., Doria et al., Reference Doria, Jaramillo and Herrera2008; Wing et al., Reference Wing, Herrera, Jaramillo, Gómez-Navarro, Wilf and Labandeira2009; Correa et al., Reference Correa, Jaramillo, Manchester and Gutierrez2010; Carvalho et al., Reference Carvalho, Herrera, Jaramillo, Wing and Callejas2011; Martínez et al., Reference Martínez, Carvalho, Madriñán and Jaramillo2015; Herrera et al., Reference Herrera, Carvalho, Wing, Jaramillo and Herendeen2019). The Paleocene floras represent the world’s oldest true Neotropical rainforests, based on family-level composition, leaf traits, and isotopic indicators for closed-canopy environments (Wing et al., Reference Wing, Herrera, Jaramillo, Gómez-Navarro, Wilf and Labandeira2009; Graham et al., Reference Graham, Herrera, Jaramillo, Wing and Freeman2019). The large fossil legumes and palm fruits are also typical of tropical forests (Gómez-Navarro et al., Reference Gómez-Navarro, Jaramillo, Herrera, Wing and Callejas2009; Herrera et al., Reference Herrera, Carvalho, Wing, Jaramillo and Herendeen2019). The three fossil sites are not as close to the KPgE in age as the others discussed here but far closer than any other tropical localities. Carvalho et al. (Reference Carvalho, Jaramillo, de la Parra, Caballero-Rodríguez, Herrera, Wing, Turner, D’Apolito, Romero-Báez, Narváez, Martínez, Gutierrez, Labandeira, Bayona, Rueda, Paez-Reyes, Cárdenas, Duque, Crowley, Santos and Silvestro2021) recently merged the macrofloral data with a finely resolved palynological dataset from Colombia covering the 72–58 Ma interval, integrating 2,053 Maastrichtian macrofossils, 4,898 Paleocene macrofossils, and 53,029 pollen occurrences of 1,048 taxa from 39 outcrop and well sections. The age estimates rely on well-established biostratigraphic correlations, and preservation of the K–Pg event horizon in one of the well sections provides a definitive constraint for pre- and post-extinction palynofloras (de la Parra et al., Reference de la Parra, Jaramillo, Kaskes, Goderis, Claeys, Villasante-Marcos, Bayona, Hatsukawa and Caballero2022). Notably, the Colombian locations are much closer to Chicxulub (ca. 2000 vs. 3,000 km) than the areas discussed in the western USA and would have frozen during the impact winter according to climate models (Figure 1).

The Carvalho et al. (Reference Carvalho, Jaramillo, de la Parra, Caballero-Rodríguez, Herrera, Wing, Turner, D’Apolito, Romero-Báez, Narváez, Martínez, Gutierrez, Labandeira, Bayona, Rueda, Paez-Reyes, Cárdenas, Duque, Crowley, Santos and Silvestro2021) analysis produced several significant outcomes, most prominently the direct linkage of the KPgE with the emergence of the first tropical rainforests. Pollen data indicated an extinction peak of ca. 45% at 66 Ma, one of the highest recorded globally, significant increases in angiosperm abundance at the expense of gymnosperms, a clear cluster separation of Cretaceous and Paleocene composition, and a ca. 6 Myr recovery of diversity. We note that these palynological results are based on composite stratigraphy, and therefore the extinction estimates are not fully comparable with single, continuous sections. The floristic shift is also striking in the macrofloras. The Guaduas flora contains a generalized mixture of plant families, including Rhamnaceae, Dilleniaceae, and Zingiberales. However, the Paleocene floras have plant-family composition and relative abundance impressively similar to the modern day, dominated by legumes and palms with Melastomataceae, Menispermaceae, Euphorbiaceae, and Malvaceae among the supporting taxa (Figure 3A). The findings from Colombia provide new, detailed evidence that the rise of angiosperms to true ecological dominance, shaping terrestrial biodiversity in the last phase of the Angiosperm Terrestrial Revolution (see Benton et al., Reference Benton, Wilf and Sauquet2022), was a direct legacy of the KPgE and the opportunities it created. The results mesh well with a state-of-the-art molecular clock study that used 238 well-constrained calibration fossils (Ramírez-Barahona et al., Reference Ramírez-Barahona, Sauquet and Magallón2020). Those authors found that angiosperms achieved much of their phylogenetic diversity as stem lineages in the Cretaceous, and crown-group families mostly evolved after the KPgE (Figure 3B).

Patagonia

Until recently, detailed knowledge of the K–Pg plant extinction in the Southern Hemisphere was mainly limited to palynological work in New Zealand, which showed a fern-spore spike associated with an event horizon and iridium anomaly, ecological disruption, and minimal overall extinction (Vajda et al., Reference Vajda, Raine and Hollis2001; Vajda and Raine, Reference Vajda and Raine2003). This situation changed with a series of investigations on Maastrichtian and Danian coastal lowland macrofloras from Chubut, Patagonian Argentina (paleolatitude ca. 50°S); these assemblages were intensively collected and well dated using combinations of radiometric, paleomagnetic, and biostratigraphic constraints (Barreda et al., Reference Barreda, Cúneo, Wilf, Currano, Scasso and Brinkhuis2012; Scasso et al., Reference Scasso, Aberhan, Ruiz, Weidemeyer, Medina and Kiessling2012; Clyde et al., Reference Clyde, Wilf, Iglesias, Slingerland, Barnum, Bijl, Bralower, Brinkhuis, Comer, Huber, Ibañez-Mejia, Jicha, Krause, Schueth, Singer, Raigemborn, Schmitz, Sluijs and Zamaloa2014; Vellekoop et al., Reference Vellekoop, Holwerda, Prámparo, Willmott, Schouten, Cúneo, Scasso and Brinkhuis2017; Clyde et al., Reference Clyde, Krause, De Benedetti, Ramezani, Cúneo, Gandolfo, Haber, Whelan and Smith2021). More than 5,000 specimens were collected from the latest Maastrichtian portion of the Lefipán Formation (ca. 66.5–66 Ma) and the Danian Salamanca (ca. 65–64 Ma) and Las Flores (ca. 62 Ma) formations, as well as smaller but systematically informative samples from the Maastrichtian-Danian La Colonia Formation (e.g., Iglesias et al., Reference Iglesias, Wilf, Johnson, Zamuner, Cúneo, Matheos and Singer2007, Reference Iglesias, Wilf, Stiles and Wilf2021; Cúneo et al., Reference Cúneo, Gandolfo, Zamaloa and Hermsen2014). The K–Pg event bed is not preserved in these strata, but the stratigraphic interval that brackets the event or its hiatus is constrained to only a few meters of section in the Lefipán and La Colonia study areas (Barreda et al., Reference Barreda, Cúneo, Wilf, Currano, Scasso and Brinkhuis2012; Clyde et al., Reference Clyde, Krause, De Benedetti, Ramezani, Cúneo, Gandolfo, Haber, Whelan and Smith2021).

The Patagonian floras have been well-studied systematically. The Lefipán and La Colonia assemblages include Araceae, lotuses, diverse aquatic ferns, and several conifer families (Cúneo et al., Reference Cúneo, Gandolfo, Zamaloa and Hermsen2014; Wilf et al., Reference Wilf, Donovan, Cúneo and Gandolfo2017; Andruchow-Colombo et al., Reference Andruchow-Colombo, Escapa, Cúneo and Gandolfo2018, Reference Andruchow‐Colombo, Gandolfo, Escapa and Cúneo2022). The Salamanca Formation and correlative strata uniquely preserve diverse fossil reproductive structures from the earliest Cenozoic of the Southern Hemisphere (Figure 4), providing extraordinary direct evidence of likely KPgE survivor taxa. Examples include multiple organs of Agathis, cocosoid palm fruits (Attaleinae), Rhamnaceae flowers (probable ziziphoid clade), two genera of Cunoniaceae flowers (Schizomerieae and an unknown clade of the family crown-group), and a Menispermaceae (Stephania) endocarp (Futey et al., Reference Futey, Gandolfo, Zamaloa, Cúneo and Cladera2012; Jud et al., Reference Jud, Gandolfo, Iglesias and Wilf2017; Escapa et al., Reference Escapa, Iglesias, Wilf, Catalano, Caraballo-Ortiz and Rubén Cúneo2018; Jud et al., Reference Jud, Gandolfo, Iglesias and Wilf2018a,Reference Jud, Iglesias, Wilf and Gandolfob; Jud and Gandolfo, Reference Jud and Gandolfo2021). Fossil leaves from the Salamanca include species of Podocarpaceae (Dacrycarpus and an extinct scale-leaved genus), Akaniaceae, Anacardiaceae, Fabaceae, Lauraceae, Nothofagaceae, and Rosaceae (Quiroga et al., Reference Quiroga, Mathiasen, Iglesias, Mill and Premoli2016; Andruchow-Colombo et al., Reference Andruchow-Colombo, Escapa, Carpenter, Hill, Iglesias, Abarzua and Wilf2019; Iglesias et al., Reference Iglesias, Wilf, Stiles and Wilf2021). Most of these taxa continued to be prominent elements of later Patagonian fossil assemblages and living Gondwana-derived rainforest floras.

Palynological analysis of the Lefipán, which includes early Danian pollen, provided the only detailed view of K–Pg plant turnover in Patagonia in one continuous section (Barreda et al., Reference Barreda, Cúneo, Wilf, Currano, Scasso and Brinkhuis2012). The results included a transient disruption followed by the recovery of nearly all Cretaceous taxa, broadly similar to prior findings from New Zealand (Vajda et al., 2001), and an intriguing spike in the earliest Danian, not of ferns but of the extinct conifer family Cheirolepidiaceae (Classopollis pollen). The palynology from all the Patagonian formations shows very similar composition (e.g., Barreda et al., Reference Barreda, Cúneo, Wilf, Currano, Scasso and Brinkhuis2012; Clyde et al., Reference Clyde, Wilf, Iglesias, Slingerland, Barnum, Bijl, Bralower, Brinkhuis, Comer, Huber, Ibañez-Mejia, Jicha, Krause, Schueth, Singer, Raigemborn, Schmitz, Sluijs and Zamaloa2014; Stiles et al., Reference Stiles, Wilf, Iglesias, Gandolfo and Cúneo2020; Clyde et al., Reference Clyde, Krause, De Benedetti, Ramezani, Cúneo, Gandolfo, Haber, Whelan and Smith2021), indicating minimal variation in higher plant taxa (as typically represented by palynomorphs; e.g., Nichols and Johnson, Reference Nichols and Johnson2008) across the K–Pg or among basins.

An early study of fossil leaves indicated that the Danian Salamanca leaf floras were more diverse than the most comparable Paleocene floras from the western USA (Iglesias et al., Reference Iglesias, Wilf, Johnson, Zamuner, Cúneo, Matheos and Singer2007). Later, Stiles et al. (Reference Stiles, Wilf, Iglesias, Gandolfo and Cúneo2020) produced the first quantitative analysis of macrofloral turnover in the Southern Hemisphere, based on the Lefipán and Salamanca dicot-leaf assemblages (n > 3,500 leaves). The authors found a sharp K–Pg drop in rarefied species diversity consistent with an extinction event (Figure 5A). Notably, both the Lefipán and Salamanca samples were more speciose than comparable North American samples (Figure 5A). Only five ‘survivor’ species were found, suggesting a > 90% face-value extinction far exceeding the North Dakota baseline of 57%. This comparison and the overall decrease in diversity (Figure 5A) strongly support significant species losses but illustrate the importance of age control when comparing extinction estimates. Given its larger age uncertainty, the Lefipán sample could correspond to any of several time slices of the Hell Creek Formation that, if taken alone for strict comparability, would also yield 90% or comparable species extinction (Figure 2B). Morphospace analysis of all leaf species unexpectedly showed that the Danian floras retained and expanded on Cretaceous morphologies, particularly by exhibiting new lobed and toothed leaf forms (Figure 5B). Because the Salamanca floras appear to primarily contain paleo-endemic taxa (Iglesias et al., Reference Iglesias, Wilf, Stiles and Wilf2021), this result supported in situ adaptation and evolution in response to an overall cooling trend (Stiles et al., Reference Stiles, Wilf, Iglesias, Gandolfo and Cúneo2020).

Figure 5. Changes in K/Pg leaf diversity and leaf morphospace in Patagonia, modified from Stiles et al. (Reference Stiles, Wilf, Iglesias, Gandolfo and Cúneo2020). (A) Rarefied richness (with 95% confidence intervals) for leaf assemblages from the latest Maastrichtian Lefipán, early Danian Salamanca (P Loros 1 and P Loros 2), and late Danian Las Flores formations from Patagonia, compared with three sampling levels from the Williston Basin in North Dakota denoted by their stratigraphic positions relative to the K/Pg event bed (also see Figure 2A). Note the sharp drop in richness in both areas and the higher richness in both time periods for the Patagonian floras. (B) Morphospace plots for Maastrichtian and Danian leaf species in Patagonia, using principal coordinates analysis (PCoA). The Danian leaves have higher morphological diversity, despite lower species richness. Icons and gray lines indicate lobed Danian outlier species and their plot locations.

Discussion

The paleobotanical data reviewed here have greatly expanded the understanding of geographic heterogeneity in the K–Pg plant extinction, while illustrating how variations in temporal, spatial, and taxonomic resolution, among other factors, limit comparisons and insights. Despite the advances from the Americas, there are very few well-dated, well-preserved macrofloras of appropriate age elsewhere (Nichols and Johnson, Reference Nichols and Johnson2008). For example, the most suitable assemblage in Europe is probably the ca. 60–61 Ma Menat Fossil Pit in Puy-de-Dôme, France, which preserves notably diverse plants and insect-feeding damage (Wappler et al., Reference Wappler, Currano, Wilf, Rust and Labandeira2009). Much remains to be learned to better trace the evolutionary and biogeographic legacies of the KPgE in the living world flora, and well-constrained paleobotanical data from new areas remain fundamental.

Maritime and latitudinal buffering of impact-winter temperatures may be top-level variables for KPgE severity (Figure 1). Most plant species are frost-intolerant, and the Earth’s surface was largely frost-free at the time (Scotese, Reference Scotese2021), suggesting that the terminal Cretaceous vegetation was highly vulnerable to the impact winter. The potential freezing of the tropics is particularly critical for understanding the dramatic turnovers observed in Colombia. The faster recoveries in the southern vs. northern Rockies directly support buffering, whereas the minimal palynological extinctions in Patagonia and New Zealand could reflect both buffering and distance from the crater. The contrasting palynological and macrofloral data from the same strata in Patagonia support the idea (see Introduction) that the K–Pg plant extinction primarily occurred at lower taxonomic levels. Conversely, Colombia’s ca. 45% pollen extinction could indicate both drastic species losses and significant elimination of higher taxa.

The end-Cretaceous plant extinction has been questioned because of the apparent lack of significant global losses at the family or major-clade level (Cascales-Miñana and Cleal, Reference Cascales-Miñana and Cleal2014; Sauquet and Magallón, Reference Sauquet and Magallón2018; Thompson and Ramírez-Barahona, Reference Thompson and Ramírez-Barahona2023). However, we hold that the K–Pg event included a massive extinction of plants and more. Estimates of extinction rates based on phylogenies of living taxa are highly uncertain (Louca and Pennell, Reference Louca and Pennell2020), particularly towards deeper nodes (O’Meara and Beaulieu, Reference O’Meara and Beaulieu2021). Attempts to quantify the KPgE should rely on the fossil record. Even though fossil data are probably insufficient to address this question globally and above the species level, every regional macrofossil study with large sample sizes and well-constrained stratigraphy shows significant species losses (usually >50%; Figures 2A, 5A). Species losses are also extinctions, and species conservation is the goal of most modern conservation efforts. Given the wide distribution and high species diversity of angiosperms by the end of the Cretaceous, significant family-level extinctions were probably statistically unlikely. This situation parallels the even more diverse insects, which also show no global extinctions at the family level in large temporal bins bracketing the KPgE (Labandeira and Sepkoski, Reference Labandeira and Sepkoski1993). However, the disappearances of specialized insect-damage morphotypes on many of the same floras discussed here suggest widespread herbivore losses (Labandeira et al., Reference Labandeira, Johnson and Wilf2002; Donovan et al., Reference Donovan, Wilf, Labandeira, Johnson and Peppe2014, Reference Donovan, Iglesias, Wilf, Labandeira and Cúneo2017). The K–Pg event was more than a mass extinction for land plants because it instantaneously transformed terrestrial ecosystems, initiating developments of paramount importance for understanding and conserving today’s biotas. These included the radiations and biogeographic movements of the survivor lineages from the KPgE to the present, novel plant–animal interactions, extensive crown-group angiosperm radiations, and the rise of Neotropical and Gondwanan rainforests.

The future

The KPgE is probably the most relevant deep-time analog for the projected modern-day extinctions, which are also occurring nearly instantaneously in geologic time. Although the worst outcomes are still avoidable (Dinerstein et al., Reference Dinerstein, Joshi, Vynne, Lee, Pharand-Deschênes, França, Fernando, Birch, Burkart, Asner and Olson2020), the coming centuries are projected to bring devastating species and ecosystem losses (Barnosky et al., Reference Barnosky, Matzke, Tomiya, Wogan, Swartz, Quental, Marshall, McGuire, Lindsey, Maguire, Mersey and Ferrer2011; Armstrong McKay et al., Reference Armstrong McKay, Staal, Abrams, Winkelmann, Sakschewski, Loriani, Fetzer, Cornell, Rockström and Lenton2022); these will surely qualify as a mass extinction, even if few plant families disappear. The predictions, although challenging to compare with fossil data, are nevertheless on a scale similar to that of the regional paleobotanical studies reviewed here. For example, a meta-analysis of conservation databases found that extinction currently threatens 39.4% of plant species (Nic Lughadha et al., Reference Nic Lughadha, Bachman, Leão, Forest, Halley, Moat, Acedo, Bacon, Brewer, Gâteblé, Gonçalves, Govaerts, Hollingsworth, Krisai‐Greilhuber, de Lirio, Moore, Negrão, Onana, Rajaovelona, Razanajatovo, Reich, Richards, Rivers, Cooper, Iganci, Lewis, Smidt, Antonelli, Mueller and Walker2020). If the projected extinctions become a reality, new evolutionary radiations and ecosystems will eventually transform the biosphere as in the geologic past, but most likely far too late to benefit humans.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/ext.2023.13.

Data availability statement

All data discussed are available in the cited literature.

Acknowledgements

We thank Julia Brugger, Santiago Ramírez- Barahona, and Nathan Jud for kindly contributing graphics; the editors and reviewers, including John Alroy, Haijun Song, Vivi Vajda, and one anonymous colleague, for encouragement and helpful comments; and Gabriella Rossetto-Harris, Hervé Sauquet, and Michael Benton for discussions.

Author contribution

Conceptual development: P.W., M.R.C., E.S; Writing—original draft: P.W.; Writing—review and editing: P.W., M.R.C., E.S.

Financial support

Recent research reviewed here that involved the authors was supported by the National Science Foundation (P.W., Grant Numbers DEB-1556666 and EAR-1925755; M.R.C., Grant Number EAR-1829299), the Smithsonian Tropical Research Institute (M.R.C.), the Geological Society of America (E.S. and M.R.C.), and the Paleontological Society (E.S.).

Competing interest

The authors declare none.

References

Alvarez, LW, Alvarez, W, Asaro, F and Michel, HV (1980) Extraterrestrial cause for the Cretaceous-Tertiary extinction: Experimental results and theoretical interpretation. Science 208, 10951108.CrossRefGoogle Scholar
Andruchow-Colombo, A, Escapa, IH, Carpenter, RJ, Hill, RS, Iglesias, A, Abarzua, AM and Wilf, P (2019) Oldest record of the scale-leaved clade of Podocarpaceae, early Paleocene of Patagonia, Argentina. Alcheringa 43, 127145.CrossRefGoogle Scholar
Andruchow-Colombo, A, Escapa, IH, Cúneo, NR and Gandolfo, MA (2018) Araucaria lefipanensis (Araucariaceae), a new species with dimorphic leaves from the Late Cretaceous of Patagonia, Argentina. American Journal of Botany 105, 10671087.CrossRefGoogle ScholarPubMed
Andruchow‐Colombo, A, Gandolfo, MA, Escapa, IH and Cúneo, NR (2022) New genus of Cupressaceae from the upper Cretaceous of Patagonia (Argentina) fills a gap in the evolution of the ovuliferous complex in the family. Journal of Systematics and Evolution 60, 14171439.CrossRefGoogle Scholar
Armstrong McKay, DI, Staal, A, Abrams, JF, Winkelmann, R, Sakschewski, B, Loriani, S, Fetzer, I, Cornell, SE, Rockström, J and Lenton, TM (2022) Exceeding 1.5°C global warming could trigger multiple climate tipping points. Science 377, eabn7950.CrossRefGoogle ScholarPubMed
Barclay, RS, Johnson, KR, Betterton, WJ and Dilcher, DL (2003) Stratigraphy and megaflora of a K-T boundary section in the eastern Denver Basin, Colorado. Rocky Mountain Geology 38, 4571.CrossRefGoogle Scholar
Bardeen, CG, Garcia, RR, Toon, OB and Conley, AJ (2017) On transient climate change at the Cretaceous−Paleogene boundary due to atmospheric soot injections. Proceedings of the National Academy of Sciences USA 114, E7415E7424.CrossRefGoogle ScholarPubMed
Barnosky, AD, Matzke, N, Tomiya, S, Wogan, GOU, Swartz, B, Quental, TB, Marshall, C, McGuire, JL, Lindsey, EL, Maguire, KC, Mersey, B and Ferrer, EA (2011) Has the Earth’s sixth mass extinction already arrived? Nature 471, 5157.CrossRefGoogle ScholarPubMed
Barreda, VD, Cúneo, NR, Wilf, P, Currano, ED, Scasso, RA and Brinkhuis, H (2012) Cretaceous/Paleogene floral turnover in Patagonia: Drop in diversity, low extinction, and a Classopollis spike. PLoS One 7, e52455.CrossRefGoogle Scholar
Benton, MJ, Wilf, P and Sauquet, H (2022) The Angiosperm Terrestrial Revolution and the origins of modern biodiversity. New Phytologist 233, 20172035.CrossRefGoogle ScholarPubMed
Berry, K (2019) Linking fern foliage with spores at the K–Pg boundary section in the Sugarite coal zone, New Mexico, USA, while questioning the orthodoxy of the global pattern of plant succession across the K–Pg boundary. Neues Jahrbuch für Geologie und Paläontologie - Abhandlungen 291, 159169.CrossRefGoogle Scholar
Berry, K (2020) Seed traits linked to differential survival of plants during the Cretaceous/Paleogene impact winter. Acta Palaeobotanica 60, 307322.CrossRefGoogle Scholar
Berry, K (2023) Fern macroflora identified in the basal Danian of the Raton formation supports palynological signal of a widespread “postdisaster” fern flora. International Journal of Plant Sciences 184, 122134.CrossRefGoogle Scholar
Brown, RW (1962) Paleocene flora of the Rocky Mountains and Great Plains. U. S. Geological Survey Professional Paper 375, 1119.Google Scholar
Brugger, J, Feulner, G, Hofmann, M and Petri, S (2021) A pronounced spike in ocean productivity triggered by the Chicxulub impact. Geophysical Research Letters 48, e2020GL092260.CrossRefGoogle Scholar
Brugger, J, Feulner, G and Petri, S (2017) Baby, it’s cold outside: Climate model simulations of the effects of the asteroid impact at the end of the Cretaceous. Geophysical Research Letters 44, 419427.CrossRefGoogle Scholar
Butrim, MJ, Royer, DL, Miller, IM, Dechesne, M, Neu-Yagle, N, Lyson, TR, Johnson, KR and Barclay, RS (2022) No consistent shift in leaf dry mass per area across the Cretaceous-Paleogene boundary. Frontiers in Plant Science 13, 894690.CrossRefGoogle ScholarPubMed
Carvalho, MR, Herrera, FA, Jaramillo, CA, Wing, SL and Callejas, R (2011) Paleocene Malvaceae from northern South America and their biogeographical implications. American Journal of Botany 98, 13371355.CrossRefGoogle ScholarPubMed
Carvalho, MR, Jaramillo, C, de la Parra, F, Caballero-Rodríguez, D, Herrera, F, Wing, S, Turner, BL, D’Apolito, C, Romero-Báez, M, Narváez, P, Martínez, C, Gutierrez, M, Labandeira, C, Bayona, G, Rueda, M, Paez-Reyes, M, Cárdenas, D, Duque, Á, Crowley, JL, Santos, C and Silvestro, D (2021) Extinction at the end-Cretaceous and the origin of modern Neotropical rainforests. Science 372, 6368.CrossRefGoogle ScholarPubMed
Cascales-Miñana, B and Cleal, CJ (2014) The plant fossil record reflects just two great extinction events. Terra Nova 26, 195200.CrossRefGoogle Scholar
Clyde, WC, Krause, JM, De Benedetti, F, Ramezani, J, Cúneo, NR, Gandolfo, MA, Haber, P, Whelan, C and Smith, T (2021) New South American record of the Cretaceous–Paleogene boundary interval (La Colonia Formation, Patagonia, Argentina). Cretaceous Research 125, 104889.CrossRefGoogle Scholar
Clyde, WC, Ramezani, J, Johnson, KR, Bowring, SA and Jones, MM (2016) Direct high-precision U–Pb geochronology of the end-Cretaceous extinction and calibration of Paleocene astronomical timescales. Earth and Planetary Science Letters 452, 272280.CrossRefGoogle Scholar
Clyde, WC, Wilf, P, Iglesias, A, Slingerland, RL, Barnum, T, Bijl, PK, Bralower, TJ, Brinkhuis, H, Comer, EE, Huber, BT, Ibañez-Mejia, M, Jicha, BR, Krause, JM, Schueth, JD, Singer, BS, Raigemborn, MS, Schmitz, MD, Sluijs, A and Zamaloa, MC (2014) New age constraints for the Salamanca formation and lower Río Chico group in the western san Jorge Basin, Patagonia, Argentina: Implications for K/Pg extinction recovery and land mammal age correlations. Geological Society of America Bulletin 126, 289306.CrossRefGoogle Scholar
Correa, E, Jaramillo, C, Manchester, S and Gutierrez, M (2010) A fruit and leaves of rhamnaceous affinities from the Late Cretaceous (Maastrichtian) of Colombia. American Journal of Botany 97, 7179.CrossRefGoogle ScholarPubMed
Cúneo, NR, Gandolfo, MA, Zamaloa, MC and Hermsen, EJ (2014) Late Cretaceous aquatic plant world in Patagonia, Argentina. PLoS One 9, e104749.CrossRefGoogle ScholarPubMed
de la Parra, F, Jaramillo, C, Kaskes, P, Goderis, S, Claeys, P, Villasante-Marcos, V, Bayona, G, Hatsukawa, Y and Caballero, D (2022) Unraveling the record of a tropical continental Cretaceous-Paleogene boundary in northern Colombia, South America. Journal of South American Earth Sciences 114, 103717.CrossRefGoogle Scholar
Dinerstein, E, Joshi, AR, Vynne, C, Lee, ATL, Pharand-Deschênes, F, França, M, Fernando, S, Birch, T, Burkart, K, Asner, GP and Olson, DA (2020) A “global safety net” to reverse biodiversity loss and stabilize Earth’s climate. Science Advances 6, eabb2824.CrossRefGoogle Scholar
Donovan, MP, Iglesias, A, Wilf, P, Labandeira, CC and Cúneo, NR (2017) Rapid recovery of Patagonian plant-insect associations after the end-Cretaceous extinction. Nature Ecology & Evolution 1, 0012.CrossRefGoogle Scholar
Donovan, MP, Wilf, P, Labandeira, CC, Johnson, KR and Peppe, DJ (2014) Novel insect leaf-mining after the end-Cretaceous extinction and the demise of Cretaceous leaf miners, Great Plains, USA. PLoS One 9, e103542.CrossRefGoogle ScholarPubMed
Dorf, E (1940) Relationship between floras of type Lance and Fort Union formations. Geological Society of America Bulletin 51, 213235.CrossRefGoogle Scholar
Doria, G, Jaramillo, CA and Herrera, F (2008) Menispermaceae from the Cerrejón Formation, middle to late Paleocene, Colombia. American Journal of Botany 95, 954973.CrossRefGoogle ScholarPubMed
Ellis, B, Johnson, KR and Dunn, RE (2003) Evidence for an in situ early Paleocene rainforest from Castle Rock, Colorado. Rocky Mountain Geology 38, 73100.CrossRefGoogle Scholar
Erdei, B, Coiro, M, Miller, I, Johnson Kirk, R, Griffith, MP and Murphy, V (2019) First cycad seedling foliage from the fossil record and inferences for the Cenozoic evolution of cycads. Biology Letters 15, 20190114.CrossRefGoogle ScholarPubMed
Escapa, IH, Iglesias, A, Wilf, P, Catalano, SA, Caraballo-Ortiz, MA and Rubén Cúneo, N (2018) Agathis trees of Patagonia’s Cretaceous-Paleogene death landscapes and their evolutionary significance. American Journal of Botany 105, 13451368.CrossRefGoogle ScholarPubMed
Flynn, AG and Peppe, DJ (2019) Early Paleocene tropical forest from the Ojo Alamo Sandstone, San Juan Basin, New Mexico, USA. Paleobiology 45, 612635.CrossRefGoogle Scholar
Futey, M, Gandolfo, MA, Zamaloa, MC, Cúneo, NR and Cladera, G (2012) Arecaceae fossil fruits from the Paleocene of Patagonia, Argentina. Botanical Review 78, 205234.CrossRefGoogle Scholar
Gómez-Navarro, C, Jaramillo, C, Herrera, F, Wing, SL and Callejas, R (2009) Palms (Arecaceae) from a Paleocene rainforest of northern Colombia. American Journal of Botany 96, 13001312.CrossRefGoogle ScholarPubMed
Gradstein, FM, Ogg, JG, Schmitz, M and Ogg, G (2012) The Geologic Time Scale 2012. Amsterdam: Elsevier.Google Scholar
Graham, HV, Herrera, F, Jaramillo, C, Wing, SL and Freeman, KH (2019) Canopy structure in Late Cretaceous and Paleocene forests as reconstructed from carbon isotope analyses of fossil leaves. Geology 47, 977981.CrossRefGoogle Scholar
Gulick, SPS, Bralower, TJ, Ormö, J, Hall, B, Grice, K, Schaefer, B, Lyons, S, Freeman, KH, Morgan, JV, Artemieva, N, Kaskes, P, de Graaff, SJ, Whalen, MT, Collins, GS, Tikoo, SM, Verhagen, C, Christeson, GL, Claeys, P, Coolen, MJL, Goderis, S, Goto, K, Grieve, RAF, McCall, N, Osinski, GR, Rae, ASP, Riller, U, Smit, J, Vajda, V, Wittmann, A and Expedition 364 Scientists (2019) The first day of the Cenozoic. Proceedings of the National Academy of Sciences USA 116, 1934219351.CrossRefGoogle ScholarPubMed
Herrera, F, Carvalho, MR, Wing, SL, Jaramillo, C and Herendeen, PS (2019) Middle to late Paleocene Leguminosae fruits and leaves from Colombia. Australian Systematic Botany 32, 385408.CrossRefGoogle Scholar
Hickey, LJ (1977) Stratigraphy and paleobotany of the Golden Valley Formation (early Tertiary) of western North Dakota. Geological Society of America Memoir 150, 1183.Google Scholar
Hotton, CL (2002) Palynology of the Cretaceous-Tertiary boundary in Central Montana: Evidence for extraterrestrial impact as a cause of the terminal Cretaceous extinctions. Geological Society of America Special Papers 361, 473501.Google Scholar
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
Iglesias, A, Wilf, P, Johnson, KR, Zamuner, AB, Cúneo, NR, Matheos, SD and Singer, BS (2007) A Paleocene lowland macroflora from Patagonia reveals significantly greater richness than North American analogs. Geology 35, 947950.CrossRefGoogle Scholar
Iglesias, A, Wilf, P, Stiles, E and Wilf, R (2021) Patagonia’s diverse but homogeneous early Paleocene forests: Angiosperm leaves from the Danian Salamanca and Peñas Coloradas formations, San Jorge Basin, Chubut, Argentina. Palaeontologia Electronica 24, a02.Google Scholar
Jaramillo, C and Cárdenas, A (2013) Global warming and Neotropical rainforests: A historical perspective. Annual Review of Earth and Planetary Sciences 41, 741766.CrossRefGoogle Scholar
Johnson, KR (2002) The megaflora of the Hell Creek and lower Fort Union formations in the western Dakotas: Vegetational response to climate change, the Cretaceous-Tertiary boundary event, and rapid marine transgression. Geological Society of America Special Papers 361, 329391.Google Scholar
Johnson, KR and Ellis, B (2002) A tropical rainforest in Colorado 1.4 million years after the Cretaceous-Tertiary boundary. Science 296, 23792383.CrossRefGoogle ScholarPubMed
Johnson, KR, Nichols, DJ, Attrep, M and Orth, CJ (1989) High-resolution leaf-fossil record spanning the Cretaceous-Tertiary boundary. Nature 340, 708711.CrossRefGoogle Scholar
Johnson, KR, Reynolds, ML, Werth, KW and Thomasson, JR (2003) Overview of the Late Cretaceous, early Paleocene, and early Eocene megafloras of the Denver Basin, Colorado. Rocky Mountain Geology 38, 101120.CrossRefGoogle Scholar
Jones, HL, Westerhold, T, Birch, H, Hull, P, Hédi Negra, M, Röhl, U, Sepúlveda, J, Vellekoop, J, Whiteside, JH, Alegret, L, Henehan, M, Robinson, L, van Dijk, J and Bralower, T (2023) Stratigraphy of the Cretaceous/Paleogene (K/Pg) boundary at the Global Stratotype Section and Point (GSSP) in El Kef, Tunisia: New insights from the El Kef Coring Project. Geological Society of America Bulletin. https://doi.org/10.1130/B36487.1.CrossRefGoogle Scholar
Jud, NA and Gandolfo, MA (2021) Fossil evidence from South America for the diversification of Cunoniaceae by the earliest Palaeocene. Annals of Botany 127, 305315.CrossRefGoogle ScholarPubMed
Jud, NA, Gandolfo, MA, Iglesias, A and Wilf, P (2017) Flowering after disaster: Early Danian buckthorn (Rhamnaceae) flowers and leaves from Patagonia. PLoS One 12, e0176164.CrossRefGoogle ScholarPubMed
Jud, NA, Gandolfo, MA, Iglesias, A and Wilf, P (2018a) Fossil flowers from the early Palaeocene of Patagonia, Argentina, with affinity to Schizomerieae (Cunoniaceae). Annals of Botany 121, 431442.CrossRefGoogle ScholarPubMed
Jud, NA, Iglesias, A, Wilf, P and Gandolfo, MA (2018b) Fossil moonseeds from the Paleogene of West Gondwana (Patagonia, Argentina). American Journal of Botany 105, 927942.CrossRefGoogle ScholarPubMed
Kiessling, W, Smith, JA and Raja, NB (2023) Improving the relevance of paleontology to climate change policy. Proceedings of the National Academy of Sciences USA 120, e2201926119.CrossRefGoogle ScholarPubMed
Koenen, EJM, Ojeda, DI, Bakker, FT, Wieringa, JJ, Kidner, C, Hardy, OJ, Pennington, RT, Herendeen, PS, Bruneau, A and Hughes, CE (2021) The origin of the legumes is a complex paleopolyploid phylogenomic tangle closely associated with the Cretaceous–Paleogene (K–Pg) mass extinction event. Systematic Biology 70, 508526.CrossRefGoogle ScholarPubMed
Labandeira, CC, Johnson, KR and Wilf, P (2002) Impact of the terminal Cretaceous event on plant-insect associations. Proceedings of the National Academy of Sciences USA 99, 20612066.CrossRefGoogle ScholarPubMed
Labandeira, CC and Sepkoski, JJ (1993) Insect diversity in the fossil record. Science 261, 310315.CrossRefGoogle ScholarPubMed
Lohaus, R and Van de Peer, Y (2016) Of dups and dinos: Evolution at the K/Pg boundary. Current Opinion in Plant Biology 30, 6269.CrossRefGoogle ScholarPubMed
Louca, S and Pennell, MW (2020) Extant timetrees are consistent with a myriad of diversification histories. Nature 580, 502505.CrossRefGoogle ScholarPubMed
Lyons, SL, Karp, AT, Bralower, TJ, Grice, K, Schaefer, B, Gulick, SPS, Morgan, JV and Freeman, KH (2020) Organic matter from the Chicxulub crater exacerbated the K–Pg impact winter. Proceedings of the National Academy of Sciences USA 117, 2532725334.CrossRefGoogle ScholarPubMed
Lyson, TR, Miller, IM, Bercovici, AD, Weissenburger, K, Fuentes, AJ, Clyde, WC, Hagadorn, JW, Butrim, MJ, Johnson, KR, Fleming, RF, Barclay, RS, Maccracken, SA, Lloyd, B, Wilson, GP, Krause, DW and Chester, SGB (2019) Exceptional continental record of biotic recovery after the Cretaceous–Paleogene mass extinction. Science 366, eaay2268.CrossRefGoogle ScholarPubMed
Manchester, SR (2014) Revisions to Roland Brown’s North American Paleocene flora. Acta Musei Nationalis Pragae 70, 153210.CrossRefGoogle Scholar
Marshall, CR (2023) Forty years later: The status of the “Big Five” mass extinctions. Cambridge Prisms: Extinction 1, 125.Google Scholar
Martínez, C, Carvalho, MR, Madriñán, S and Jaramillo, CA (2015) A late Cretaceous Piper (Piperaceae) from Colombia and diversification patterns for the genus. American Journal of Botany 102, 273289.CrossRefGoogle ScholarPubMed
McElwain, JC and Punyasena, SW (2007) Mass extinction events and the plant fossil record. Trends in Ecology & Evolution 22, 548557.CrossRefGoogle ScholarPubMed
McLoughlin, S, Carpenter, RJ and Pott, C (2011) Ptilophyllum muelleri (Ettingsh.) comb. nov. from the Oligocene of Australia: Last of the Bennettitales? International Journal of Plant Sciences 172, 574585.CrossRefGoogle Scholar
Merkhofer, L, Wilf, P, Haas, MT, Kooyman, RM, Sack, L, Scoffoni, C and Cúneo, NR (2015) Resolving Australian analogs for an Eocene Patagonian paleorainforest using leaf size and floristics. American Journal of Botany 102, 11601173.CrossRefGoogle ScholarPubMed
Morgan, JV, Bralower, TJ, Brugger, J and Wünnemann, K (2022) The Chicxulub impact and its environmental consequences. Nature Reviews Earth & Environment 3, 338354.CrossRefGoogle Scholar
Nic Lughadha, E, Bachman, SP, Leão, TCC, Forest, F, Halley, JM, Moat, J, Acedo, C, Bacon, KL, Brewer, RFA, Gâteblé, G, Gonçalves, SC, Govaerts, R, Hollingsworth, PM, Krisai‐Greilhuber, I, de Lirio, EJ, Moore, PGP, Negrão, R, Onana, JM, Rajaovelona, LR, Razanajatovo, H, Reich, PB, Richards, SL, Rivers, MC, Cooper, A, Iganci, J, Lewis, GP, Smidt, EC, Antonelli, A, Mueller, GM and Walker, BE (2020) Extinction risk and threats to plants and fungi. Plants, People, Planet 2, 389408.CrossRefGoogle Scholar
Nichols, DJ and Fleming, RF (2002) Palynology and palynostratigraphy of Maastrichtian, Paleocene, and Eocene strata in the Denver Basin, Colorado. Rocky Mountain Geology 37, 135163.CrossRefGoogle Scholar
Nichols, DJ and Johnson, KR (2008) Plants and the K-T Boundary. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
O’Meara, B and Beaulieu, J (2021) Potential survival of some, but not all, diversification methods. EcoEvoRxiv, https://doi.org/10.32942/osf.io/w5nvd.Google ScholarPubMed
Pillay, R, Venter, M, Aragon‐Osejo, J, González‐Del‐Pliego, P, Hansen, AJ, Watson, JE and Venter, O (2022) Tropical forests are home to over half of the world’s vertebrate species. Frontiers in Ecology and the Environment 20, 1015.CrossRefGoogle ScholarPubMed
Quiroga, MP, Mathiasen, P, Iglesias, A, Mill, RR and Premoli, AC (2016) Molecular and fossil evidence disentangle the biogeographical history of Podocarpus, a key genus in plant geography. Journal of Biogeography 43, 372383.CrossRefGoogle Scholar
Ramírez-Barahona, S, Sauquet, H and Magallón, S (2020) The delayed and geographically heterogeneous diversification of flowering plant families. Nature Ecology & Evolution 4, 12321238.CrossRefGoogle ScholarPubMed
Raup, DM and Sepkoski, JJ (1982) Mass extinctions in the marine fossil record. Science 215, 15011503.CrossRefGoogle ScholarPubMed
Raynolds, RG, Johnson, KR, Ellis, B, Dechesne, M and Miller, IM (2007) Earth history along Colorado’s front range: Salvaging geologic data in the suburbs and sharing it with the citizens. GSA Today 17, 410.CrossRefGoogle Scholar
Sauquet, H, Ho, SY, Gandolfo, MA, Jordan, GJ, Wilf, P, Cantrill, DJ, Bayly, MJ, Bromham, L, Brown, GK, Carpenter, RJ, Lee, DM, Murphy, DJ, Sniderman, JM and Udovicic, F (2012) Testing the impact of calibration on molecular divergence times using a fossil-rich group: The case of Nothofagus (Fagales). Systematic Biology 61, 289313.CrossRefGoogle Scholar
Sauquet, H and Magallón, S (2018) Key questions and challenges in angiosperm macroevolution. New Phytologist 219, 11701187.CrossRefGoogle ScholarPubMed
Scasso, RA, Aberhan, M, Ruiz, L, Weidemeyer, S, Medina, FA and Kiessling, W (2012) Integrated bio- and lithofacies analysis of coarse-grained, tide-dominated deltaic environments across the Cretaceous/Paleogene boundary in Patagonia, Argentina. Cretaceous Research 36, 3757.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, SP, 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, Whalen, MT and Willumsen, PS (2010) The Chicxulub asteroid impact and mass extinction at the Cretaceous-Paleogene boundary. Science 327, 12141218.CrossRefGoogle ScholarPubMed
Scotese, CR (2021) An atlas of Phanerozoic paleogeographic maps: The seas come in and the seas go out. Annual Review of Earth and Planetary Sciences 49, 679728.CrossRefGoogle Scholar
Slik, JWF, Arroyo-Rodríguez, V, Aiba, S-I, Alvarez-Loayza, P, Alves, LF, Ashton, P, Balvanera, P, Bastian, ML, Bellingham, PJ, van den Berg, E, Bernacci, L, da Conceição Bispo, P, Blanc, L, Böhning-Gaese, K, Boeckx, P, Bongers, F, Boyle, B, Bradford, M, Brearley, FQ, Breuer-Ndoundou Hockemba, M, Bunyavejchewin, S, Calderado Leal Matos, D, Castillo-Santiago, M, Catharino, ELM, Chai, SL, Chen, Y, Colwell, RK, Chazdon, RL, Clark, C, Clark, DB, Clark, DA, Culmsee, H, Damas, K, Dattaraja, HS, Dauby, G, Davidar, P, DeWalt, SJ, Doucet, JL, Duque, A, Durigan, G, Eichhorn, KAO, Eisenlohr, PV, Eler, E, Ewango, C, Farwig, N, Feeley, KJ, Ferreira, L, Field, R, de Oliveira Filho, AT, Fletcher, C, Forshed, O, Franco, G, Fredriksson, G, Gillespie, T, Gillet, JF, Amarnath, G, Griffith, DM, Grogan, J, Gunatilleke, N, Harris, D, Harrison, R, Hector, A, Homeier, J, Imai, N, Itoh, A, Jansen, PA, Joly, CA, de Jong, BHJ, Kartawinata, K, Kearsley, E, Kelly, DL, Kenfack, D, Kessler, M, Kitayama, K, Kooyman, R, Larney, E, Laumonier, Y, Laurance, S, Laurance, WF, Lawes, MJ, Amaral, IL, Letcher, SG, Lindsell, J, Lu, X, Mansor, A, Marjokorpi, A, Martin, EH, Meilby, H, Melo, FPL, Metcalfe, DJ, Medjibe, VP, Metzger, JP, Millet, J, Mohandass, D, Montero, JC, de Morisson Valeriano, M, Mugerwa, B, Nagamasu, H, Nilus, R, Ochoa-Gaona, S, Onrizal Page, N, Parolin, P, Parren, M, Parthasarathy, N, Paudel, E, Permana, A, Piedade, MTF, Pitman, NCA, Poorter, L, Poulsen, AD, Poulsen, J, Powers, J, Prasad, RC, Puyravaud, JP, Razafimahaimodison, JC, Reitsma, J, dos Santos, JR, Roberto Spironello, W, Romero-Saltos, H, Rovero, F, Rozak, AH, Ruokolainen, K, Rutishauser, E, Saiter, F, Saner, P, Santos, BA, Santos, F, Sarker, SK, Satdichanh, M, Schmitt, CB, Schöngart, J, Schulze, M, Suganuma, MS, Sheil, D, da Silva Pinheiro, E, Sist, P, Stevart, T, Sukumar, R, Sun, IF, Sunderland, T, Suresh, HS, Suzuki, E, Tabarelli, M, Tang, J, Targhetta, N, Theilade, I, Thomas, DW, Tchouto, P, Hurtado, J, Valencia, R, van Valkenburg, JLCH, van do, T, Vasquez, R, Verbeeck, H, Adekunle, V, Vieira, SA, Webb, CO, Whitfeld, T, Wich, SA, Williams, J, Wittmann, F, Wöll, H, Yang, X, Adou Yao, CY, Yap, SL, Yoneda, T, Zahawi, RA, Zakaria, R, Zang, R, de Assis, RL, Garcia Luize, B and Venticinque, EM (2015) An estimate of the number of tropical tree species. Proceedings of the National Academy of Sciences USA 112, 74727477.CrossRefGoogle ScholarPubMed
Stiles, E, Wilf, P, Iglesias, A, Gandolfo, MA and Cúneo, NR (2020) Cretaceous–Paleogene plant extinction and recovery in Patagonia. Paleobiology 46, 445469.CrossRefGoogle Scholar
Thompson, JB and Ramírez-Barahona, S (2023) No evidence for angiosperm mass extinction at the Cretaceous–Paleogene (K–Pg) boundary. bioRxiv https://doi.org/10.1101/2023.02.15.528726.Google Scholar
Traverse, A (1988) Plant evolution dances to a different beat. Historical Biology 1, 277301.CrossRefGoogle Scholar
Tschudy, RH, Pillmore, CL, Orth, CJ, Gilmore, JS and Knight, JD (1984) Disruption of the terrestrial plant ecosystem at the Cretaceous-Tertiary boundary, Western Interior. Science 225, 10301032.CrossRefGoogle ScholarPubMed
Vajda, V and Bercovici, A (2014) The global vegetation pattern across the Cretaceous–Paleogene mass extinction interval: A template for other extinction events. Global and Planetary Change 122, 2949.CrossRefGoogle Scholar
Vajda, V, Ocampo, A, Ferrow, E and Koch, CB (2015) Nano particles as the primary cause for long-term sunlight suppression at high southern latitudes following the Chicxulub impact— Evidence from ejecta deposits in Belize and Mexico. Gondwana Research 27, 10791088.CrossRefGoogle Scholar
Vajda, V and Raine, JI (2003) Pollen and spores in marine Cretaceous/Tertiary boundary sediments at mid-Waipara River, North Canterbury, New Zealand. New Zealand Journal of Geology and Geophysics 46, 255273.CrossRefGoogle Scholar
Vajda, V, Raine, JI and Hollis, CJ (2001) Indication of global deforestation at the Cretaceous-Tertiary boundary by New Zealand fern spike. Science 294, 17001702.CrossRefGoogle ScholarPubMed
Vanneste, K, Baele, G, Maere, S and Van De Peer, Y (2014) Analysis of 41 plant genomes supports a wave of successful genome duplications in association with the Cretaceous–Paleogene boundary. Genome Research 24, 13341347.CrossRefGoogle ScholarPubMed
Vellekoop, J, Holwerda, F, Prámparo, MB, Willmott, V, Schouten, S, Cúneo, NR, Scasso, RA and Brinkhuis, H (2017) Climate and sea-level changes across a shallow marine Cretaceous-Palaeogene boundary succession in Patagonia, Argentina. Palaeontology 60, 519534.CrossRefGoogle Scholar
Wake, DB and Vredenburg, VT (2008) Are we in the midst of the sixth mass extinction? A view from the world of amphibians. Proceedings of the National Academy of Sciences USA 105, 1146611473.CrossRefGoogle ScholarPubMed
Wappler, T, Currano, ED, Wilf, P, Rust, J and Labandeira, CC (2009) No post-Cretaceous ecosystem depression in European forests? Rich insect-feeding damage on diverse middle Palaeocene plants at Menat, France. Proceedings of the Royal Society B 276, 42714277.CrossRefGoogle ScholarPubMed
Wilf, P (2008) Fossil angiosperm leaves: Paleobotany’s difficult children prove themselves. Paleontological Society Papers 14, 319333.CrossRefGoogle Scholar
Wilf, P, Cúneo, NR, Escapa, IH, Pol, D and Woodburne, MO (2013) Splendid and seldom isolated: The paleobiogeography of Patagonia. Annual Review of Earth and Planetary Sciences 41, 561603.CrossRefGoogle Scholar
Wilf, P, Donovan, MP, Cúneo, NR and Gandolfo, MA (2017) The fossil flip-leaves (Retrophyllum, Podocarpaceae) of southern South America. American Journal of Botany 104, 13441369.CrossRefGoogle ScholarPubMed
Wilf, P and Escapa, IH (2016) Molecular dates require geologic testing. New Phytologist 209, 13591362.CrossRefGoogle ScholarPubMed
Wilf, P and Johnson, KR (2004) Land plant extinction at the end of the Cretaceous: A quantitative analysis of the North Dakota megafloral record. Paleobiology 30, 347368.2.0.CO;2>CrossRefGoogle Scholar
Wilson, PK, Wilson Mantilla, GP and Strӧmberg, CAE (2021) Seafood salad: A diverse latest Cretaceous florule from eastern Montana. Cretaceous Research 121, 104734.CrossRefGoogle Scholar
Wing, SL (2004) Mass extinctions in plant evolution. In Taylor, PD (ed.), Extinctions in the History of Life. Cambridge, UK: Cambridge University Press, pp. 6197.CrossRefGoogle Scholar
Wing, SL, Alroy, J and Hickey, LJ (1995) Plant and mammal diversity in the Paleocene to early Eocene of the Bighorn Basin. Palaeogeography, Palaeoclimatology, Palaeoecology 115, 117155.CrossRefGoogle Scholar
Wing, SL and Currano, ED (2013) Plant response to a global greenhouse event 56 million years ago. American Journal of Botany 100, 12341254.CrossRefGoogle ScholarPubMed
Wing, SL, Herrera, F, Jaramillo, CA, Gómez-Navarro, C, Wilf, P and Labandeira, CC (2009) Late Paleocene fossils from the Cerrejón Formation, Colombia, are the earliest record of Neotropical rainforest. Proceedings of the National Academy of Sciences USA 106, 1862718632.CrossRefGoogle ScholarPubMed
Wolfe, JA (1987) Late Cretaceous-Cenozoic history of deciduousness and the terminal Cretaceous event. Paleobiology 13, 215226.CrossRefGoogle Scholar
Wolfe, JA and Upchurch, GR (1986) Vegetation, climatic and floral changes at the Cretaceous-Tertiary boundary. Nature 324, 148152.CrossRefGoogle Scholar
Wright, IJ, Dong, N, Maire, V, Prentice, IC, Westoby, M, Díaz, S, Gallagher, RV, Jacobs, BF, Kooyman, R, Law, EA, Leishman, MR, Niinemets, Ü, Reich, PB, Sack, L, Villar, R, Wang, H and Wilf, P (2017) Global climatic drivers of leaf size. Science 357, 917921.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. Modeled surface air temperatures (A) before and (B) during the coldest impact-winter year, 3 years after the Chicxulub impact; pickaxe icons indicate the principal fossil floras discussed in the text from (north to south) the northern Great Plains, southern Rockies, Colombia, and southern Argentina. Drafted by J. Brugger, using model output from Brugger et al. (2021), with permission.

Figure 1

Figure 2. Plant extinction in the Williston Basin, North Dakota. (A) Ranges in the composite section of 141 macrofossil species occurring in more than one stratigraphic bin, with 99% confidence intervals on last appearances, relative to the K/Pg event horizon recognized from multiple indicators. The conservative estimate of 57% species extinction is based on this analysis, using only the last appearances in the uppermost 5 m of Cretaceous strata (Wilf and Johnson, 2004). The observed extinction increases if the window is shifted down, as demonstrated in (B), showing simulated extinctions and species richness from the same dataset (singletons included) when based on discrete 10-m bin windows (Stiles et al., 2020). Modified from (A) Wilf and Johnson (2004) and (B) Stiles et al. (2020).

Figure 2

Figure 3. The post-KPgE rise of extant angiosperm families, seen in Colombian macrofossils and a molecular-phylogenetic reconstruction of stem- vs. crown-group age distributions. In (A), bars show proportions of species (with standard deviations) in a family in a compilation of 72 Neotropical vegetation plots; darker bars indicate the top-10 families that account for half of total stems. Red diamonds mark families present in a latest Cretaceous flora; blue circles show families recognized in two Paleocene floras. Modified from Carvalho et al. (2021). (B) Frequency distributions and 95% kernel density curves of reconstructed mean stem (red) and crown (blue) angiosperm family ages, from an analysis calibrated using 238 well-constrained fossils. Subscript “m” indicates age estimates derived from molecular analyses rather than direct geological dating. Modified from Ramírez-Barahona et al. (2020), with permission.

Figure 3

Figure 4. Early Danian reproductive fossils from the Salamanca Formation, Chubut, Argentina. (A) Agathis immortalis (Araucariaceae) branch with leaves and axillary pollen cones (see Escapa et al., 2018). (B) Notiantha grandensis (Rhamnaceae) flowers side-by-side in transverse and longitudinal views (see Jud et al., 2017). (C) Stephania psittaca (Menispermaceae) endocarp (see Jud et al., 2018b). Lacinipetalum spectabilum (Cunoniaceae) flower, detail of ovary surface covered in trichomes and entrapped pollen grains, glowing under epifluorescence (see Jud et al., 2018a). All the specimens are housed at Museo Paleontolόgico Egidio Feruglio, Trelew, Chubut. Credits: A, C, D: P. Wilf; B, N. Jud, used with permission.

Figure 4

Figure 5. Changes in K/Pg leaf diversity and leaf morphospace in Patagonia, modified from Stiles et al. (2020). (A) Rarefied richness (with 95% confidence intervals) for leaf assemblages from the latest Maastrichtian Lefipán, early Danian Salamanca (P Loros 1 and P Loros 2), and late Danian Las Flores formations from Patagonia, compared with three sampling levels from the Williston Basin in North Dakota denoted by their stratigraphic positions relative to the K/Pg event bed (also see Figure 2A). Note the sharp drop in richness in both areas and the higher richness in both time periods for the Patagonian floras. (B) Morphospace plots for Maastrichtian and Danian leaf species in Patagonia, using principal coordinates analysis (PCoA). The Danian leaves have higher morphological diversity, despite lower species richness. Icons and gray lines indicate lobed Danian outlier species and their plot locations.

Author comment: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R0/PR1

Comments

Dear John, Barry, and the team,

Thank you very much for inviting this submission and checking an early draft for suitability. Fortunately, we have been able to complete the submission much earlier than anticipated. We hope you find it really interesting and suitable for the series! Please note that I will be out of the country March 12-30 with little or no email access. Best, Peter

Review: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: The manuscript by Wilf et al. provides an interesting overview of published paleobotantical records across the K-Pg boundary in North and South America .The results highlight heterogeneity among records, potentially reflecting a heterogenous response the the K-Pg boundary perturbations.

Overall, the manuscript is well-written, albeit with a somewhat unconventional structure. An example is the unexpected side-discussion on geological records versus molecular clock data (lines 180-192) that is placed a bit awkwardly.

Perhaps the main take-home message is that the paleobotantical records across the K-Pg boundary are actually still rather centered around the Americas, making it difficult to truly assess a global record across the K-Pg boundary. This point could be stressed a bit more in the conclusions.

A few other comments and suggestions:

Lines 75-76:

While the KPME was in all likelihood indeed more rapid than any other geological mass extinction, I would be cautious to use the term ‘severity’ in this sentence, as other mass extinctions (e.g. PT boundary) appear to have resulted in higher overall extinction rates.

Lines 105-106:

I would refrain from using such terminology as “any KPME for plants” and “the plant KPME”, since a mass extinction event is per definition an extinction event across many different biological groups. Hence, one should not say “a mass extinction among plants”, but rather “an extinction event among plants”, of, if you will, “a massive extinction among plants…across the KPg boundary”

Please be careful with this terminology throughout the manuscript.

Line 143:

“70,000 years” gives the suggestion of a level of accuracy that is not really there. Even when adding the “ca.” to this number, it is better to use something like “ca. 70 kyrs”. The alternative option would be to add quantitative uncertainties on this number, but I don’t believe those are available either.

Lines 332-334:

There are a few other interesting sites in Europe, for example the mid-Seelandinian floras of the Gelinden Marls in Belgium, already described in the 19th century by De Saporta et al. and others, where there recently also have been studies on the plant-insect interactions (Zambon et al, 2023).

Lines 374-375: see earlier comments on the terminology around mass extinctions. Please refrain from using terminology such as “a mass extinction for…” followed by a specific biological group.

All in all, these minor points leads me to suggest this manuscript to be accepted for publication with minor revisions.

Review: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R0/PR3

Conflict of interest statement

No competing Interests

Comments

Comments to Author: This is an interesting and useful overview of the vegetation change and mass extinction related to the end-Cretaceous event 66 million years ago. The text is very well written, easy to follow and keeps the reader interested. It has the North American record in the focus but provides a good global overview supported by figures and diagrams, also including climate effects and the future aspect on the 6th mass extinction. I suggest publication with minor revision.

1. I find the usage of the abbreviation KPME completely unacceptable. P stands for the Permian and should definitely not be used for the Paleogene. If the authors find the need to abbreviate the Cretaceous-Paleogene mass extinction then I suggest K-PgME.

In some places, the authors have used KTME, mainly in the Fig. captions. That is less bad but probably not correct either – please search through the manuscript for these inconsistencies.

2. I lack a short paragraph on holdover taxa in refugia in general. There is data from e.g. Tasmania with plant groups surviving the K-PgME until the Eocene.

3. In the section concerning the impact winter, carbon particles are mentioned as nucleus for aerosols. Importantly also nano-particles of Fe (material dispersed from the asteroid) has shown to play a major role for the following darkness and cooling. https://doi.org/10.1016/j.gr.2014.05.009.

I look forward to see this interesting paper published.

Sincerely

Vivi Vajda

Recommendation: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R0/PR4

Comments

Comments to Author: Dear Dr. Wilf,

Thank you for submitting your manuscript (ID: EXT-22-0048) to Cambridge Prisms: Extinction. I have now received reports from two reviewers. Collectively, both Reviewers think this paper is well written and significant and should be published. However, as you will see from the reports, the reviewers raised some minor issues that need to be revised or reconsidered, e.g., the abbreviation KPME. Reviewer 1 suggested to raise aspects that need further interpretation, such as the fact that the paleobotanical record across the K-PG boundary is still actually more centered on the Americas, making it difficult to truly assess the global record across the K-Pg boundary. Considering these comments, my decision is a minor revision.

Please make sure all points raised by the reviewers are addressed and provide a point-by-point response to these comments along with your revision.

Thank you for your understanding and I look forward to receiving the revised manuscript.

Sincerely,

Haijun

Prof. Haijun Song

Handling Editor, Cambridge Prisms: Extinction

Email: haijunsong@cug.edu.cn

Decision: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R0/PR5

Comments

No accompanying comment.

Author comment: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R1/PR6

Comments

Please see the response letter uploaded at Step 1 as the cover letter.

→NOTES ON FIGURE UPLOADS: I re-uploaded all five figure files here because they were not showing in the generated proof. In addition, I converted Fig4 from tif to jpeg (minimum compression) because the system kept corrupting the tif file. Fig. 4A shows as too blue in the system previews but has correct color temperature in the uploaded file. The figure files are otherwise unchanged from the initial submission.

Review: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R1/PR7

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: Thank you for addressing all the issues. I look forward to see the paper in print.

Regards

Vivi Vajda

Recommendation: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R1/PR8

Comments

Comments to Author: Dear Dr. Wilf,

Thank you again for submitting your manuscript (ID: EXT-22-0048.R1) to Cambridge Prisms: Extinction. Your revisions adequately address the concerns raised in the review and I am pleased to confirm that your paper “The end-Cretaceous plant extinction: heterogeneity, ecosystem transformation, and insights for the future” has been accepted for publication in Cambridge Prisms: Extinction.

Yours sincerely,

Haijun

Sincerely,

Haijun

Prof. Haijun Song

Handling Editor, Cambridge Prisms: Extinction

Email: haijunsong@cug.edu.cn

Decision: The end-Cretaceous plant extinction: Heterogeneity, ecosystem transformation, and insights for the future — R1/PR9

Comments

No accompanying comment.