Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-27T19:16:35.687Z Has data issue: false hasContentIssue false

Causes and consequences of end-Ediacaran extinction: An update

Published online by Cambridge University Press:  15 May 2023

Simon A.F. Darroch*
Affiliation:
Department of Earth and Environmental Sciences, Vanderbilt University, Nashville, TN, USA Evolutionary Studies Institute, Vanderbilt University, Nashville, TN, USA Senckenberg Research Institute and Museum of Natural History, Frankfurt, Germany
Emily F. Smith
Affiliation:
Department of Earth and Planetary Sciences, Johns Hopkins University, Baltimore, MD, USA
Lyle L. Nelson
Affiliation:
Department of Earth Sciences, Carleton University, Ottawa, ON, Canada
Matthew Craffey
Affiliation:
Department of Biological Sciences, University of Nebraska, Lincoln, NE, USA
James D. Schiffbauer
Affiliation:
Department of Geological Sciences, University of Missouri, Columbia, MO, USA X-Ray Microanalysis Laboratory, University of Missouri, Columbia, MO, USA
Marc Laflamme
Affiliation:
Department of Chemical and Physical Sciences, University of Toronto Mississauga, Mississauga, ON, Canada
*
Corresponding author: S.A.F. Darroch; Email: simon.a.darroch@vanderbilt.edu
Rights & Permissions [Opens in a new window]

Abstract

Since the 1980s, the existence of one or more extinction events in the late Ediacaran has been the subject of debate. Discussion surrounding these events has intensified in the last decade, in concert with efforts to understand drivers of global change over the Ediacaran–Cambrian transition and the appearance of the more modern-looking Phanerozoic biosphere. In this paper we review the history of thought and work surrounding late Ediacaran extinctions, with a particular focus on the last 5 years of paleontological, geochemical, and geochronological research. We consider the extent to which key questions have been answered, and pose new questions which will help to characterize drivers of environmental and biotic change. A key challenge for future work will be the calculation of extinction intensities that account for limited sampling, the duration of Ediacaran ‘assemblage’ zones, and the preponderance of taxa restricted to a single ‘assemblage’; without these data, the extent to which Ediacaran bioevents represent genuine mass extinctions comparable to the ‘Big 5’ extinctions of the Phanerozoic remains to be rigorously tested. Lastly, we propose a revised model for drivers of late Ediacaran extinction pulses that builds off recent data and growing consensus within the field. This model is speculative, but does frame testable hypotheses that can be targeted in the next decade of work.

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 majority of extinction-based paleontological research over the last four decades has focused on the ‘big 5’ mass extinctions of the Phanerozoic. In parallel, however, geologists and paleontologists working in the Precambrian have mulled the existence of one or more pulses of extinction (and potentially ‘mass extinction’) in the latest Neoproterozoic (~574–539 million years ago) shortly before the onset of the Cambrian. These episodes of global biotic turnover removed the mysterious Ediacara biota, as well as groups of more recognizable animal fossils. In this review, we summarize the history of ideas and research surrounding these events, as well as recent work in a range of fields that is attempting to identify the drivers – both biotic and abiotic – of extinction. We outline four key questions which, we argue, will help us to compare the causes and consequences of Ediacaran extinction alongside the Phanerozoic ‘Big 5’, and which will help us decide whether the ‘Big 5’ might eventually become the ‘Big 6’ (or the ‘Big 7’, if the current biodiversity crisis is considered). Finally, we propose a model for drivers of late Ediacaran extinction that builds off recent data. This model is speculative, but frames testable hypotheses that will help determine the role these events may have played in the Ediacaran–Cambrian emergence of the modern-looking biosphere, and thus the extent to which Ediacaran extinction and the Cambrian explosion may be linked.

Introduction

The Ediacaran–Cambrian (E–C) transition arguably marks the most important geobiological revolution of the past billion years, characterized by large perturbations to global geochemical cycles, a permanent step-change in the character of the sedimentary record, the rise of macroscopic eukaryotic life, and potentially one or more pulses of mass extinction. Although all aspects of this interval have been the subjects of intense research efforts over the last three decades, the existence of putative biotic turnover events in the latest Ediacaran has received particular attention. These events may not only have played a crucial role in fueling evolutionary radiation during the ‘rise of animals’ and acted as a powerful influence on the appearance of metazoan ecosystems with a more modern-looking structure (Knoll and Carroll, Reference Knoll and Carroll1999; Droser et al., Reference Droser, Tarhan and Gehling2017; Darroch et al., Reference Darroch, Smith, Laflamme and Erwin2018a,Reference Darroch, Laflamme and Wagnerb), but they also have invited comparison with the ‘Big 5’ mass extinctions of the Phanerozoic (Raup and Sepkoski, Reference Raup and Sepkoski1982), and thus may yield more general lessons about the causes and consequences of these catastrophic events in Earth’s history.

Five years ago, Darroch et al., Reference Darroch, Smith, Laflamme and Erwin2018a, summarized the evidence for one or more pulses of extinction in the late Ediacaran and presented a series of key questions that would be crucial for driving knowledge forward in this field. Since then many of these questions have been explored, with new paleontological, geochemical, and geochronological datasets providing the scaffolding required for building our understanding of this interval. In this review we summarize recent work surrounding the end-Ediacaran extinction events, examine the extent to which the questions posed in 2018 have been answered, and propose new questions, challenges, and research avenues that will continue to illuminate the changes that occurred over the E–C transition.

Ediacaran fossils and early animals

The late Ediacaran is characterized by the presence of macroscopic body fossils that are typically categorized as belonging to one of two faunas: either (1) ‘Ediacara biota’ – an enigmatic collection of soft-bodied organisms with uncertain relationships to extant animal phyla, and which have been subdivided into morphogroups (see, e.g., Erwin et al., Reference Erwin, Laflamme, Tweedt, Sperling, Pisani and Peterson2011; Laflamme et al., Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013); or (2) true metazoans – referring to fossils that can be more readily allied with living animal groups. However, this subdivision within Ediacaran organisms is becoming increasingly more obsolete with recent developmental (Gold et al., Reference Gold, Runnegar, Gehling and Jacobs2015), and phylogenetic (Dunn et al., Reference Dunn, Liu, Grazhdankin, Vixseboxse, Flanerry-Sutherland, Green, Harris, Wilby and Donoghue2021) data suggesting that many representatives of the Ediacara biota are likely stem-group members of known eumetazoan clades. On the other hand, given that these two different categories of organisms appear and disappear at different times in the fossil record (see, e.g., Figure 1) and possess strong morphological differences (including, for example, the presence/absence of a body plan that is present among extant phyla), classifying Ediacaran-aged taxa as ‘Ediacara biota’ vs. ‘metazoans’ is arguably still useful, and provides a heuristic model with which to explore their faunal dynamics. So, while we refer to ‘Ediacara biota’ and metazoans over the course of this review, we emphasize that this does not preclude members of the Ediacara biota as belonging to animal clades.

Figure 1. Updated summary figure illustrating the stratigraphic distribution and diversity among groups of Ediacara biota, as well as metazoans, bilaterian ichnogenera, and a δ13C curve (compiled from Yang et al. (Reference Yang, Rooney, Condon, Li, Grazhdankin, Bowyer, Hu, Macdonald and Zhu2021), Bowyer et al. (Reference Bowyer, Zhuravlev, Wood, Shields, Zhou, Curtis, Poulton, Condon, Yang and Zhu2022), and references therein). The stratigraphic ranges of the Pentaradialomorpha and Tetraradialomorpha are currently uncertain, but currently constrained by a detrital zircon age of 556 ± 24 Ma obtained from the Bonney Sandstone in South Australia (Ireland et al., Reference Ireland, Flöttmann, Fanning, Gibson and Preiss1998). Solid colors represent minimum age estimates (where available), while shaded regions represent uncertain range estimates where taxa are found beneath (or between) dated horizons. Extinction intensities – as percentage of genera lost – are given for the two putative extinction pulses at the White Sea-Nama and the E–C boundaries; intensities were calculated by simply measuring the proportion of surviving genera over total genera in the preceding assemblage zone (although see discussion in the text surrounding problems with calculating these transition).

History

The history of thought surrounding the existence of end-Ediacaran extinction events is closely linked to work defining, dating, and characterizing the base of the Cambrian. Early attempts to define this boundary were spearheaded by decades of dedicated stratigraphic, paleontological, and geochronological studies conducted by the International Geoscience Programme (IGCP), reviewed nearly 30 years ago by Brasier et al. (Reference Brasier, Cowie and Taylor1994). Nevertheless, these efforts were overwhelmingly focused on the subsequent ‘explosion’ of animal phyla in the early Cambrian, rather than the disappearance of Ediacaran soft-bodied organisms below the boundary. This is perhaps not surprising given that, prior to propositions by Seilacher (Reference Seilacher, Holland and Trendall1984, Reference Seilacher1985, Reference Seilacher1989, Reference Seilacher1992), the Ediacara biota were overwhelmingly interpreted as belonging to extant metazoan groups (e.g., Glaessner, Reference Glaessner1984; Gehling, Reference Gehling1991). As such, the fossil record of the E–C transition could be satisfactorily explained as the result of taphonomic biases towards the preservation of biomineral shells, teeth, and bones (Gehling, Reference Gehling1999). In contrast, Seilacher (Reference Seilacher1989) was largely alone in arguing that many, if not all, Ediacaran fossils represented neither true metazoans nor their earliest stem ancestors, and therefore must have suffered an extinction at some point during the E–C transition.

The case for extinction was reinvigorated by chemostratigraphic studies that identified multiple, large carbon isotope fluctuations in the late Neoproterozoic potentially tied to major upheavals in the carbon cycle; one that particularly stands-out as potentially coeval with extinction is a negative excursion that reaches values as low as −9 ‰, known now as the basal Cambrian carbon isotope excursion (or ‘BACE’), which coincides with the E–C boundary (Kirschvink et al., Reference Kirschvink, Magaritz, Ripperdan, Zhuravlev and Rozanov1991; Knoll and Walter, Reference Knoll and Walter1992; Narbonne et al., Reference Narbonne, Kaufman and Knoll1994). In the decade that followed the “Decision on the Precambrian-Cambrian boundary stratotype” (Brasier et al., Reference Brasier, Cowie and Taylor1994), notable discussion on a transitional Ediacaran–Cambrian extinction arose. For instance, noting parallels with the Permo-Triassic boundary, Knoll and Carroll (Reference Knoll and Carroll1999) stated a clear case for a mass extinction separating the Ediacaran and Cambrian faunas – a case that was only strengthened with the recognition that the BACE event also coincided with the global and synchronous disappearance of biomineralizing fossils that characterize the latest Ediacaran (Amthor et al., Reference Amthor, Grotzinger, Shröder, Bowring, Ramezani, Martin and Matter2003). The presence of earlier extinction events, however, only became recognized with more focused biostratigraphic work, and, in particular, attempts to stratigraphically subdivide the late Ediacaran.

Waggoner (Reference Waggoner1999, Reference Waggoner2003) identified three broad communities of Ediacara biota, which are still broadly thought to represent three chronologically and environmentally distinct assemblages. From oldest to youngest these are: (1) the Avalon Assemblage (~574–558 Ma), characterized by deep-water communities (Narbonne, Reference Narbonne2005; Liu et al., Reference Liu, Kenchington and Mitchell2015); (2) the White Sea Assemblage (~558–550 Ma), which represents the apex of diversity and disparity among Ediacara biota (Grazhdankin, Reference Grazhdankin2004; Droser and Gehling, Reference Droser and Gehling2015); and (3) the Nama Assemblage (~550–538 Ma), which records a drop in the diversity of Ediacara biota, alongside an expansion in several modes of metazoan ‘ecosystem engineering’ including increased trace fossil diversity, the advent of macroscopic biomineralization, and widespread suspension feeding (Germs, Reference Germs1972; Wood and Curtis, Reference Wood and Curtis2014; Schiffbauer et al., Reference Schiffbauer, Huntley, O’Neil, Darroch, Laflamme and Cai2016; Darroch et al., Reference Darroch, Smith, Laflamme and Erwin2018a,Reference Darroch, Laflamme and Wagnerb). Due to the apparent loss in diversity among Ediacara biota, the transition from the White Sea assemblage to the Nama assemblage has also been suggested as recording an extinction event. Although discussion surrounding biotic turnover at the White Sea–Nama transition has intensified recently (e.g., Darroch et al., Reference Darroch, Smith, Laflamme and Erwin2018a,Reference Darroch, Laflamme and Wagnerb; Tarhan et al., Reference Tarhan, Droser, Cole and Gehling2018; Muscente et al., Reference Muscente, Bykova, Boag, Buatois, Mángano, Eleish, Prabhu, Pan, Meyer, Schiffbauer, Fox, Hazen and Knoll2019; Evans et al., Reference Evans, Tu, Rizzo, Suprenant, Boan, McCandless, Marshall, Xiao and Droser2022), the formerly recognized ‘Kotlinian crisis’ in the southern Urals and East European Platform, which removed diversity among Ediacara biota prior to the appearance of biomineralizing metazoans like Cloudina (Brasier, Reference Brasier1992), may be time-equivalent with this transition (Grazhdankin, Reference Grazhdankin2014).

The last 10 years have seen an abundance of work on putative late Ediacaran extinction events, focusing on the possible causes and consequences of extinction pulses, as well as to what extent extinction may be linked to the Cambrian explosion. Key work in this area is summarized below.

Late Ediacaran bioevents and extinction models

Darroch et al. (Reference Darroch, Smith, Laflamme and Erwin2018a) argued for two pulses of Ediacaran extinction: one between the White Sea and Nama assemblages at ~550 Ma, and another at the E–C boundary itself. This inference was supported by Muscente et al. (Reference Muscente, Bykova, Boag, Buatois, Mángano, Eleish, Prabhu, Pan, Meyer, Schiffbauer, Fox, Hazen and Knoll2019), who used a network analysis of fossil communities together with their associated paleoenvironments to demonstrate that turnover was unlikely to be the result of a secular facies bias (see also Evans et al., Reference Evans, Tu, Rizzo, Suprenant, Boan, McCandless, Marshall, Xiao and Droser2022). The first extinction pulse apparently removed many of the most charismatic (and enigmatic) groups of Ediacara biota best known from White Sea-aged fossil localities in South Australia and Russia – principally the dickinsoniomorphs, triradialomorphs, tetraradialomorphs, and bilateromorphs – leaving a relatively species-poor assemblage dominated by erniettomorphs, arboreomorphs, and rangeomorphs in Nama-aged strata. Interestingly, genus richness among White Sea and Nama-aged rangeomorphs is also substantially lower than from the Avalon, suggesting that, while the rangeomorphs survived the first pulse of extinction, they were nonetheless negatively impacted. In contrast, genus richness among the erniettomorphs is equivalent or even potentially higher in the Nama, suggesting a positive response to the removal of White Sea-aged shallow-marine biocoenoses, and/or to the environmental conditions that pervaded the terminal Ediacaran. Although extinction is markedly focused within specific groups of Ediacara biota, many metazoan genera present in the White Sea were also affected (Figure 1). In concert with the disappearance of Ediacara biota, the first extinction pulse is also marked by the widespread appearance of more recognizable metazoans, including increased diversity (and/or behavioral disparity) in bilaterian tracemakers (Mángano and Buatois, Reference Mángano and Buatois2014; Darroch et al., Reference Darroch, Cribb, Buatois, Germs, Kenchington, Smith, Mocke, O’Neil, Schiffbauer, Maloney, Racicot, Turk, Gibson, Almond, Koester, Boag, Tweedt and Laflamme2021), the appearance of tube-dwelling animals with debated affinities (see, e.g., Schiffbauer et al., Reference Schiffbauer, Selly, Jacquet, Merz, Nelson, Strange, Cai and Smith2020; Shore et al., Reference Shore, Wood, Curtis and Bowyer2020), calcifying and sessile lophotrochozoans (Shore et al., Reference Shore, Wood, Butler, Zhuravelev, McMahon, Curtis and Bowyer2021), and rare body fossils of segmented bilaterians plausibly representing early annelids or panarthropods (Chen et al., Reference Chen, Zhou, Yuan and Xiao2019) (Figure 2). Unlike representatives of the Ediacara biota, these organisms can be more confidently allied with groups and lineages that persisted into the Cambrian (e.g., Yang et al., Reference Yang, Steiner, Zhu, Guoxiang, Liu and Liu2016, Reference Yang, Steiner, Schiffbauer, Selly, Wu, Zhang and Liu2020). Given the apparent vermiform character of much of this Nama-aged metazoan fauna, whether preserved as body- or trace fossils, Schiffbauer et al. (Reference Schiffbauer, Huntley, O’Neil, Darroch, Laflamme and Cai2016) referred to this interval as ‘Wormworld’.

Figure 2. Putative late Ediacaran ecosystem engineers, including bilaterian tracemaking behaviors that involve sediment ‘bulldozing’ and biomixing (A - Parapsammichnites), bioirrigation (B-C - large treptichnids), and suspension feeders such as Paleophragmodictya (D-E; sp. nos. P32338 and P32332-P32352 respectively, South Australia Museum), biomineralizing Cloudina (F), and other unidentified tubefauna (G). Lastly, many Ediacara biota may have also had important ecosystem engineering impacts; the enigmatic taxa Ernietta (J), Arkarua (H; sp. no. P26768, South Australia Museum), Tribrachidium (I; sp. no. N3993/5056, Palaeontological Institute, Moscow) and Pteridinium (K) are all also thought to have functioned as suspension feeders, and thus played a crucial role in forging energetic links between the pelagic and benthic realms (Cracknell et al., Reference Cracknell, Ankor, García-Bellido, Gehling, Darroch and Rahman2021; Darroch et al., Reference Darroch, Gibson, Syversen, Rahman, Racicot, Dunn, Gutarra-Diaz, Schindler, Wehrmann and Laflamme2022). Specimens shown in A–C, F, G, J and K from the Nama Group of southern Namibia (all Urusis Fm., with the exception of Ernietta shown in J from the Dabis Fm.), and photographed in the field. Filled scale bars = 1 cm, open scale bars = 5 mm.

An apparent second extinction pulse occurs at the E–C boundary, demarcated by the disappearance of almost all remaining Ediacara biota, as well as much of the metazoan fauna that characterizes the Nama Assemblage (in particular the calcifying taxa Cloudina and Namacalathus (Amthor et al., Reference Amthor, Grotzinger, Shröder, Bowring, Ramezani, Martin and Matter2003; Smith et al., Reference Smith, Nelson, Strange, Eyster, Rowland, Schrag and Macdonald2016, as well as cosmopolitan tube-dwelling forms such as Shaanxilithes and Gaojiashania (Zhu et al., Reference Zhu, Zhuravlev, Wood, Zhao and Sukhov2017)). We note that estimates of extinction intensity – particularly with respect to tube-dwelling taxa – over this pulse is complicated by a lack of consensus in taxonomic studies, and thus to what extent latest Ediacaran and earliest Cambrian tubefauna may be related. For example, recent studies suggest that some Ediacaran-type biomineralizing taxa may persist into the early Cambrian (e.g., Zhu et al., Reference Zhu, Zhuravlev, Wood, Zhao and Sukhov2017; Yang et al., Reference Yang, Rooney, Condon, Li, Grazhdankin, Bowyer, Hu, Macdonald and Zhu2021), albeit in limited localities and numbers (Cai et al., Reference Cai, Xiao, Li and Hua2019). By way of contrast, few late Ediacaran taxa have unambiguously been identified from the Cambrian, and many that do are tentative descendants (e.g., cambroctoconids), and/or re-appear in considerably modified form (see, e.g., Park et al., Reference Park, Jung, Lee, Lee, Zhen, Hua, Warren and Hughes2021). In general, more focused systematic work on the affinities of, and relationships between, late Ediacaran and early Cambrian tubefauna is sorely needed (Schiffbauer et al., Reference Schiffbauer, Rosbach, Pulsipher, Leibach, Nolan, Tang, Lindsay-Kauffman and Selly2022).

In terms of what may have driven these two events, Darroch et al., Reference Darroch, Smith, Laflamme and Erwin2018a, discussed evidence for two hypotheses – termed ‘catastrophe’ and ‘biotic replacement’ – representing the summation of ideas and data given in previous studies (principally Amthor et al., Reference Amthor, Grotzinger, Shröder, Bowring, Ramezani, Martin and Matter2003; Erwin and Tweedt, Reference Erwin and Tweedt2012; Laflamme et al., Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013; and Darroch et al., Reference Darroch, Sperling, Boag, Racicot, Mason, Morgan, Tweedt, Myrow, Erwin and Laflamme2015). The ‘catastrophe’ model suggested that Ediacaran extinction events were driven by environmental perturbations, reflected in the negative carbon isotope excursions during the E–C transition. This model invokes parallels with several of the Phanerozoic ‘Big 5’ extinctions, in particular those coinciding with the Permian–Triassic and Triassic-Jurassic boundaries. In contrast, ‘biotic replacement’ suggested that the extinction events were instead driven by the emergence of new metazoan ecosystem engineering behaviors and their associated downstream geobiological impacts, which permanently altered marine environments in a fashion that was deleterious to soft-bodied Ediacara biota. There are fewer clear parallels for this process in the Phanerozoic; however, one hypothesized cause of the late Devonian mass extinction centers on the initial radiation of terrestrial forests, which significantly influenced weathering patterns leading to eutrophication, anoxia, and prolonged intervals of ecological stress (Algeo and Scheckler, Reference Algeo and Scheckler1998; Lu et al., Reference Lu, Lu, Ikejiri, Sun, Carroll, Blair, Algeo and Sun2021). This model, therefore, identifies the emergence of new ecosystem engineers as the ultimate driver of mass extinction, albeit through a complex series of terrestrial–marine teleconnections (e.g., Lu et al., Reference Lu, Lu, Ikejiri, Sun, Carroll, Blair, Algeo and Sun2021). Other studies have focused on the impact of humans as ecosystem engineers and as a driver of the on-going ‘6th mass extinction’ (e.g., Yeakel et al., Reference Yeakel, Pires, de Aguiar, O’Donnell, Guimaraes, Gravel and and Gross2020; Pineda-Munoz et al., Reference Pineda-Munoz, Wang, Lyons, Toth and McGuire2021), potentially raising interesting parallels between the first and most recent mass extinctions of macroscopic life. In general, the extent to which the emergence of new ecosystem engineering behaviors in deep time had led to extinction, or instead evolutionary radiations, is a question that has long required more focused work (see, e.g., Erwin, Reference Erwin2008).

Do these intervals of biotic turnover represent (‘mass’) extinctions at all?

Our understanding of the geochronology, chemostratigraphy, and biostratigraphy of the late Ediacaran has increased substantially in the last 10 years. However, given uncertainties surrounding stratigraphic correlation between sites (Xiao et al., Reference Xiao, Narbonne, Zhou, Laflamme, Grazhdankin, Moczydlowska-Vidal and Cui2016), the placement of key boundaries (Nelson et al., Reference Nelson, Ramezani, Almond, Darroch, Taylor, Brenner, Furey, Turner and Smith2022), and the mechanisms of fossil preservation (Laflamme et al., Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013; Slagter et al., Reference Slagter, Hao, Planavsky, Konhauser and Tarhan2022; Gibson et al., Reference Gibson, Schiffbauer, Wallace and Darroch2023), a reasonable question is: do these apparent intervals of biotic turnover really represent extinction intervals (or more specifically, ‘mass’ extinctions) at all? The iconic Ediacaran fossil sites in South Australia illustrate some of these issues; they are among the best-studied Ediacaran localities, are frequently taken to epitomize the diversity and community structure of ‘White Sea’-aged assemblages, and almost always make their way into analyses of Ediacaran diversity through time (see, e.g., Laflamme et al., Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013; Darroch et al., Reference Darroch, Sperling, Boag, Racicot, Mason, Morgan, Tweedt, Myrow, Erwin and Laflamme2015; Darroch et al., Reference Darroch, Laflamme and Wagner2018b; Eden et al., Reference Eden, Manica and Mitchell2022; Evans et al., Reference Evans, Tu, Rizzo, Suprenant, Boan, McCandless, Marshall, Xiao and Droser2022). However, beyond post-dating the Shuram-Wonoka isotope excursion and a UPb detrital zircon age of 556 ± 24 Ma obtained from the underlying Bonney Sandstone (Ireland et al., Reference Ireland, Flöttmann, Fanning, Gibson and Preiss1998), the age of the principal fossiliferous horizons in the Ediacara Member are unconstrained. Given that South Australia preserves a number of Ediacaran morphogroups thought to be disappeared over the White Sea-Nama transition, this is perhaps not a trivial barrier to inferring an extinction event.

Mass extinctions are typically identified as episodes of anomalously high rates of taxonomic loss, occurring on global scales, that are approximately synchronous over a relatively short interval of geological time (exactly how ‘short’ is an evolving field, but currently thought to be ~105 years; see Burgess et al., Reference Burgess, Bowring and Shen2014). We argue that the White Sea-Nama transition and E–C boundary currently satisfy one of these three criteria – specifically, being a global vs. regional signal. Older work (Laflamme et al., Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013; Boag et al., Reference Boag, Darroch and Laflamme2016) has suggested – and more recent work (Boddy et al., Reference Boddy, Mitchell, Merdith and Liu2022; Evans et al., Reference Evans, Tu, Rizzo, Suprenant, Boan, McCandless, Marshall, Xiao and Droser2022) has confirmed – that there are no obvious geographical, facies, or preservational biases that can readily explain the loss of taxa, nor account for the observation that entire morphogroups (dickinsonimorpha, bilateromorpha, etc.) are lost, and thus extinction is apparently taxonomically clustered. The number and wide geographic spread of dated Nama-aged fossil sites also provides some support for dismissing the notion that White Sea-aged communities persist and remain widespread into the Nama. With respect to the issue with South Australian fossils mentioned above, it is thus far more likely that these communities are White Sea in age (coeval with well-dated horizons in Russia), rather than a totally unique Nama-aged locality.

The question as to rates (and magnitude) of taxonomic loss is harder to address. Calculating the simple proportion of surviving genera (over total genera in the preceding assemblage zone) gives genus extinction intensities of 74.1% for the White Sea-Nama (WS–NM) transition, and 91% over the Nama-Fortunian (NM–FN) – magnitudes noted by previous studies as being comparable to those estimated for many of the ‘Big 5’ (see, e.g., Evans et al., Reference Evans, Tu, Rizzo, Suprenant, Boan, McCandless, Marshall, Xiao and Droser2022). However, raw percentages are strongly biased by variations in sampling intensity, and quantifying extinction rates over the late Ediacaran is fraught with other difficulties stemming from the character of the Ediacaran fossil record, and ongoing difficulties with stratigraphic correlation and subdivision. For example, from our occurrence dataset (Figure 1; taxa and references provided in Supplementary Material) we can argue for major extinctions across the White Sea-Nama and E–C boundaries utilizing a simple proportion of extinct genera, but we do not have much confidence in Ediacaran per capita extinction rates using common methods such as Foote (Reference Foote1999), Alroy (Reference Alroy2008), or Alroy (Reference Alroy2014). The principal issue is that these methods utilize the proportion of taxa crossing boundaries for their extinction rate estimates in various patterns (e.g., boundary crossers, 3-timers, and gap fillers spanning 3 intervals with no detections in the middle) and the Ediacaran is dominated by taxa that occur in a single assemblage zone. This likely results from limited sampling, high turnover, and the long duration of these zones (~10–15 Ma). Thus, the effective sample size for an analysis of per capita Extinction rates across the Ediacaran is very small, leading to estimates of per capita extinction at the E–C boundary that have dubious reliability (for example: 2.48 per Foote, Reference Foote1999, 3.04 per Alroy, Reference Alroy2014, utilizing the R package ‘divDyn’ Kocsis et al., Reference Kocsis, Reddin, Alroy and Kiessling2019). The most direct quantitative solution would be to examine extinction rates at finer timescales than assemblage zones, but there is – as yet – no unifying framework for subdividing these intervals, and there is limited chronostratigraphic data to applying such a framework across global collections. We can only argue, for now, for the presence of high extinction at the E–C transition based on apparent turnover patterns, but providing firm quantitative support will require additional work, and (potentially) the application of other methods for extinction rate modeling.

The last criterion – that extinction is rapid and synchronous – is, similarly hard to satisfy, and something discussed in more detail below (see the section titled ‘Key questions in 2023’).

Another recent challenge to the existence of late Ediacaran extinction events has come from phylogenetic modeling, framing biotic patterns over the E–C transition as an artifact of evolutionary patterns and stem- vs. crown-group diversity dynamics (i.e., suggesting that Ediacaran extinction and Cambrian explosion are different facets of the same process). For example, Budd and Mann (Reference Budd and Mann2020) have used birth-death models to argue that, if all Ediacara biota are viewed as stem-group members of extant bilaterian clades (and assuming a high level of background extinction), then the proportion of diversity within the total group can be quickly ‘drowned’ by the crown, thus mimicking a mass extinction. However, these models invoke a consistent rate of background extinction, and so cannot explain the synchronous and global loss of multiple morphogroups over, for example, the White Sea–Nama transition. Consequently, these models do not provide a good match for the observed diversity trends, although we note that Budd and Mann’s (Reference Budd and Mann2020) models incorporating mass extinction events do provide a match, and so may be relevant for strikingly different reasons.

Key questions in 2018

Darroch et al. (Reference Darroch, Smith, Laflamme and Erwin2018a) emphasized that the ‘catastrophe’ and ‘biotic replacement’ models were not mutually exclusive, but noted that each bring contrasting predictions, and moreover could be tested by addressing four key questions. Below, we briefly re-cap these questions (along with appropriate context), before reviewing recent work in these areas and assessing to what extent these questions have been answered.

1. What, and when, was the Shuram? In the Phanerozoic, several of the ‘Big 5’ mass extinctions are associated with perturbations to global geochemical cycles, which in turn are recorded in global isotope records (see, in particular, large igneous provinces as drivers of the Permian–Triassic (Shen et al., Reference Shen, Crowley, Wang, Bowring, Erwin, Sadler, Cao, Rothman, Henderson, Ramezani, Zhang, Shen, Wang, Wang, Mu, Li, Liu, Liu, Zheng, Y-F and Jin2011) and Triassic–Jurassic (Ruhl et al., Reference Ruhl, Bonis, Reichert, Damste and Kurschner2011) mass extinctions). Consequently, the existence of a large carbon isotope excursion in the late Neoproterozoic -– the Shuram event – has long been suspected as a potential source of environmental stress (see discussion in Tarhan et al., Reference Tarhan, Droser, Cole and Gehling2018). The Shuram is among the largest negative carbon isotope excursions in Earth history, with carbonate δ13C values as low as −12‰ recorded on multiple paleocontinents (e.g., Grotzinger et al., Reference Grotzinger, Fike and Fischer2011). It has been proposed that this excursion records massive perturbations to global geochemical cycles (e.g., Fike et al., Reference Fike, Grotzinger, Pratt and Summons2006), which could plausibly represent a source of physiological stress and a possible driving mechanism for late Ediacaran extinction. However, in 2018 there were poor radiometric age constraints for the onset, duration, and recovery from the Shuram excursion, limiting efforts to test for temporal correlation with extinction. Furthermore, there was disagreement over whether the excursion should be interpreted as a primary marine (e.g., Husson et al., Reference Husson, Higgins, Maloof and Schoene2015) or diagenetic (e.g., Knauth and Kennedy, Reference Knauth and Kennedy2009) signal.

2. What was the BACE? Similar to the Shuram, the BACE is a large negative carbon isotope excursion (δ13C values < −6‰) recorded in multiple localities worldwide. In 2018 the BACE had better age constraints than the Shuram and was recognized as coinciding with a second extinction pulse at the E–C boundary (Amthor et al., Reference Amthor, Grotzinger, Shröder, Bowring, Ramezani, Martin and Matter2003). Like the Shuram, however, there were outstanding questions as to what the excursion represented, what its precise timing and duration was, and whether it was primary or diagenetic, locally or globally controlled. All of these uncertainties limited interpretations for causal linkages between excursion and extinction.

3. Can we disentangle correlation versus causation in late Ediacaran extinction events? This question builds from the previous two in emphasizing the need for plausible cause and effect in extinction studies – a standard set by workers over the last decade on the Permian–Triassic extinction involving integrated geochronology, geochemistry, and paleontology (e.g., Shen et al., Reference Shen, Crowley, Wang, Bowring, Erwin, Sadler, Cao, Rothman, Henderson, Ramezani, Zhang, Shen, Wang, Wang, Mu, Li, Liu, Liu, Zheng, Y-F and Jin2011; Burgess et al., Reference Burgess, Bowring and Shen2014; Clarkson et al., Reference Clarkson, Kasemann, Wood, Lenton, Daines, Richoz, Ohnemueller, Meixner, Poulton and Tipper2015). In 2018, although there was potential temporal correlation when discussing both the ‘catastrophe’ and ‘biotic replacement’ models, evidence for causation was lacking. For example, with the ‘catastrophe’ model there was little idea as to what E–C carbon isotope excursions represented, precluding discussion of links between environmental change and sources of biotic stress. Likewise, with ‘biotic replacement’ there were significant knowledge gaps surrounding how the Ediacara biota and emerging metazoan fauna interacted, both as individual taxa and within communities. Consequently, there was minimal evidence for biotic interactions – antagonistic or otherwise – and thus no substantiated mechanism for a biotic driver of extinction.

4. What role did the end-Ediacaran extinction play in the Cambrian Explosion? Several of the Phanerozoic ‘Big 5’ mass extinctions were followed by radiation in surviving clades, expansions of morphologic disparity, and rapid diversification of new taxa (e.g., post-K/Pg radiations of mammals (O’Leary et al., Reference O’Leary, Bloch, Flynn, Gaudin, Giallombardo, Giannini, Goldberg, Kraatz, Luo, Meng, Ni, Novacek, Perini, Randall, Rougier, Sargis, Silcox, Simmons, Spaulding, Velazco, Weksler, Wible and Cirranello2013), birds (Ksepka et al., Reference Ksepka, Stidham and Williamson2017), and mollusks (Krug and Jablonski, Reference Krug and Jablonski2012). An intriguing question, therefore, is to what extent the Cambrian explosion could have been triggered (or perhaps driven) by late Ediacaran extinction pulses as a response to an ‘ecological vacuum’ (e.g., Knoll and Carroll, Reference Knoll and Carroll1999).

Have these questions been answered?

Questions 1 and 2 (‘what, and when, were the Shuram and BACE isotope excursions?’) have arguably received the most attention over the last five years, with new geochemical and geochronological data bringing these events into sharper focus. With respect to the Shuram, recent Re-Os dates from Northwest Canada and Oman have demonstrated that: (1) on separate paleocontinents, the excursions are synchronous within the error of these radioisotopic measurements, and (2) the excursion lasted <6.7 ± 5.6 million years from c. 574 Ma to c. 567 Ma (Rooney et al., Reference Rooney, Cantine, Bergmann, Gomez-Perez, Al Baloushi, Boag, Busch, Sperling and Strauss2020). Additional radioisotopic and chemostratigraphic data from Newfoundland are consistent with this finding and suggest that the Shuram carbon isotope excursion began after 571 Ma and ended before 562 Ma (Canfield et al., Reference Canfield, Knoll, Poulton, Narbonne and Dunning2020). These new data demonstrate that this perturbation did not coincide with the White Sea–Nama transition (at c. 550 Ma (Bowring et al., Reference Bowring, Grotzinger, Condon, Ramezani, Newall and Allen2007), as previously suggested, and, furthermore, demonstrate that it postdated both the Gaskiers glaciation and the earliest dated macrofossils of the Avalon assemblage (Macdonald et al., Reference Macdonald, Strauss, Sperling, Halverson, Narbonne, Johnston, Kunzmann, Schrag and Higgins2013; Pu et al., Reference Pu, Bowring, Ramezani, Myrow, Raub, Landing, Mills, Hodgin and Macdonald2016; Matthews et al., Reference Matthews, Liu, Yang, McIlroy, Levell and Condon2021).

While the ultimate cause of the Shuram excursion remains contentious, many have suggested that mechanisms implicate changes in marine redox conditions (e.g., Fike et al., Reference Fike, Grotzinger, Pratt and Summons2006; Zhang et al., Reference Zhang, Xiao, Romaniello, Hardisty, Li, Melezhik, Pokrovsky, Cheng, Shi, Lenton and Anbar2019; Li et al., Reference Li, Cao, Lloyd, Algeo, Zhao, Wang, Zhao and Chen2020) and/or changes in the locus of primary productivity (e.g., Busch et al., Reference Busch, Hodgin, Ahm, Husson, Macdonald, Bergmann, Higgins and Strauss2022). These changes are not mutually exclusive and could relate to external factors such as fluctuations in eustatic sea level (Busch et al., Reference Busch, Hodgin, Ahm, Husson, Macdonald, Bergmann, Higgins and Strauss2022) and/or nutrient availability (Cañadas et al., Reference Cañadas, Papineau, Leng and Li2022). Other work has challenged the interpretation that large Neoproterozoic carbon isotope excursions record the global dissolved inorganic reservoir composition, but suggest they are still coeval responses to external forcings, such as primary production and sea level changes (e.g., Ahm et al., Reference Ahm, Maloof, Macdonald, Hoffman, Bjerrum, Bold, Rose, Strauss and Higgins2019). Regardless of its origin, at present it seems unlikely that the cause of the Shuram played any role in driving late Ediacaran extinction pulses (save, perhaps, for a decline in acanthomorphic acritarch assemblages (Ouyang et al., Reference Ouyang, Zhou, Xiao, Chen and Shao2019; Yang et al., Reference Yang, Rooney, Condon, Li, Grazhdankin, Bowyer, Hu, Macdonald and Zhu2021), but may – interestingly – have played a role in driving origination.

While a link between the Shuram excursion and first pulse of extinction at the White Sea–Nama transition is irreconcilable with new radioisotopic constraints, Yang et al. (Reference Yang, Rooney, Condon, Li, Grazhdankin, Bowyer, Hu, Macdonald and Zhu2021) have provided evidence for a second negative excursion that postdates the Shuram and ended at ~550 Ma. This excursion is potentially correlative with the carbon isotope excursion documented in the basal Nama Group of southern Namibia >548 Ma (Bowring et al., Reference Bowring, Grotzinger, Condon, Ramezani, Newall and Allen2007; Wood et al., Reference Wood, Poulton, Prave, Hoffmann, Clarkson, Guilbaud, Lyne, Tostevin, Bowyer, Penny, Curtis and Kasemann2015) and/or negative carbon isotope values documented in the Stirling Quartzite of Death Valley, California (Verdel et al., Reference Verdel, Wernicke and Bowring2011). The existence of a large negative excursion coincident with the White Sea-Nama transition would invite suggestions of causality, although more data will be required to establish beyond doubt the existence of two large negative carbon isotope excursions in the middle Ediacaran. Furthermore, Eden et al. (Reference Eden, Manica and Mitchell2022) found no evidence for a ‘catastrophe’-type signature in Nama-aged communities – something that might support a causal relationship with extinction. Nevertheless, the rift-related Central Iapetus magmatic province (CIMP)—with loosely constrained pulses from c. 580–550 associated with the opening of the Iapetus Ocean (Youbi et al., Reference Youbi, Ernst, Söderlund, Boumehdi, Lahna, Tassinari, El Moume, Bensalah, Adatte, Bond and Keller2020)—provides a prospective mechanism for environmental and/or carbon isotope perturbations in the middle Ediacaran, worthy of further investigation (Figure 3).

Figure 3. Ediacaran–Cambrian rift volcanism in Laurentia and timing of interpreted rifts leading to passive margin development around Laurentia after the breakup of Rodinia. Rifting along the southwestern margin has recently been suggested to coincide with the E–C boundary, potentially resulting in the BACE negative carbon isotope excursion and a second pulse of Ediacaran extinction (see Hodgin et al., Reference Hodgin, Nelson, Wall, Barron-Diaz, Webb, Schmitz, Fike, Hagadorn and Smith2021; Smith et al., Reference Smith, Nelson, O’Connell, Eyster and Lonsdale2022). CAMP, Central Atlantic magmatic province; CIMP, Central Iapetus magmatic province; Panth, Panthalassa Ocean.

With respect to the BACE, although large, negative carbon isotope excursions have been found at horizons approximately coinciding with the E–C boundary on several paleocontinents, there remain significant challenges with correlations and temporal calibration (e.g., Bowyer et al., Reference Bowyer, Zhuravlev, Wood, Shields, Zhou, Curtis, Poulton, Condon, Yang and Zhu2022). Beyond understanding the carbon isotope record, these correlations have significant bearing on how the biostratigraphic records of the E–C boundary interval are interpreted, and thus our understanding of rates and patterns of biotic turnover at this boundary remains limited. Previous calibration of the BACE (and of the E–C boundary) at ~541 Ma hinged on a U–Pb ID-TIMS ash bed date just below onset of the negative carbon isotope excursion in the Ara Group of Oman (Bowring et al., Reference Bowring, Grotzinger, Condon, Ramezani, Newall and Allen2007). However, more recent data from Namibia, South Africa, and northern Mexico have revised this calibration (Linnemann et al., Reference Linnemann, Ovtcharova, Schaltegger, Gartner, Hautmann, Geyer, Vickers-Rich, Rich, Plessen, Hofmann, Zieger, Krause, Kriesfield and Smith2019; Hodgin et al., Reference Hodgin, Nelson, Wall, Barron-Diaz, Webb, Schmitz, Fike, Hagadorn and Smith2021; Nelson et al., Reference Nelson, Ramezani, Almond, Darroch, Taylor, Brenner, Furey, Turner and Smith2022). Nelson et al. (Reference Nelson, Ramezani, Almond, Darroch, Taylor, Brenner, Furey, Turner and Smith2022) suggest that, given available age constraints, the BACE and the last appearance of Ediacaran fossils, such as cloudinomorphs and erniettomorphs should be interpreted as <538 Ma. Regardless of its absolute age, as postulated when it was first identified in the 1990s, the BACE continues to hold up as a useful marker of the E–C boundary (e.g., Narbonne et al., Reference Narbonne, Kaufman and Knoll1994; Corsetti and Hagadorn, 2000). Furthermore, more recent data continue to reinforce that the BACE postdated most occurrences of Ediacaran fossils such as erniettomorphs and cloudinomorphs (e.g., Smith et al., Reference Smith, Nelson, O’Connell, Eyster and Lonsdale2022) and predated most occurrences of Cambrian trace fossils and small shelly fossils (e.g., Topper et al., Reference Topper, Betts, Dorjnamjaa, Li, Li, Altanshagai, Enkhbaatar and Skovsted2022), supporting hypotheses of biotic turnover across this excursion.

These issues with correlation feed into discussion of what may have caused the BACE and to what extent it represents a plausible driver for extinction. For example, if emerging geochronological data reinforce the existing difficulties with correlating the BACE across different continents, then this might suggest that it represents a regional or diagenetic signal. We note, however, that the excursion persists across regional sequence boundaries and dolomitization fronts and is thus inconsistent with a purely diagenetic origin (Smith et al., Reference Smith, Nelson, O’Connell, Eyster and Lonsdale2022). Assuming that the excursion represents a perturbation to the global marine carbon cycle, Hodgin et al. (Reference Hodgin, Nelson, Wall, Barron-Diaz, Webb, Schmitz, Fike, Hagadorn and Smith2021) suggested a genetic relationship with rift-related volcanism in southern Laurentia. This hypothesis suggests that dissolved inorganic carbon of marine waters attained highly 13C-depleted compositions due to the combined influx of mantle carbon from volcanic outgassing and combusted organic carbon within intruded sedimentary rift-basins, drawing parallels between end-Ediacaran extinction and the Permian–Triassic and Triassic-Jurassic mass extinction events. While this would be consistent with a ‘catastrophe’ model for the end-Ediacaran extinction, the proposed mechanistic link remains speculative, as it hinges on temporal correlation between the BACE and pulses of volcanism associated with the c. 539.5–530.0 Ma Wichita igneous province (Hodgin et al., Reference Hodgin, Nelson, Wall, Barron-Diaz, Webb, Schmitz, Fike, Hagadorn and Smith2021; Wall et al., 2021), as well as stratigraphic correlation to other basaltic and volcaniclastic units in southwestern North America (e.g., Smith et al., Reference Smith, Nelson, O’Connell, Eyster and Lonsdale2022; see Figure 3).

This work addressing questions 1 and 2 has naturally fed into question 3 – disentangling correlation vs. causation – in tying the late Ediacaran carbon isotope excursions to potential sources of ecological stress, and thus targeting the ‘catastrophe’ model for late Ediacaran extinction. However, substantial work has also been done in this vein targeting the ‘biotic replacement’ model, which suggests that the disappearance of the Ediacara biota was a consequence of ecosystem engineering by the emerging metazoan fauna. Two key questions in this regard have therefore been: (1) how intense was ecosystem engineering in the Ediacaran? And (2) how did the Ediacara biota and metazoan fauna interact?

In terms of the former, much recent effort has focused on the trace fossil record. Not only do bioturbating animals have powerful effects on resource flows and in modifying the physical environment (Rhoads et al., Reference Rhoads, Yingst, Ullman and Wiley1978; Jones et al., Reference Jones, Lawton and Shachak1994; Rosenberg et al., Reference Rosenberg, Nilsson and Diaz2001; Meysman et al., Reference Meysman, Middelburg and Heip2006), but they are also a record of ecosystem engineering that is relatively easy to preserve as fossils (Marenco and Bottjer, Reference Marenco, Bottjer, Cuddington, Byers, Wilson and Hastings2007). A detailed search through Nama-aged sediments in several localities worldwide has uncovered a remarkable diversity of Ediacaran ichnotaxa (e.g., Parry et al., Reference Parry, Boggiani, Condon, Garwood, Leme, McIlroy, Brasier, Trindade, Campanha, Pacheco, Diniz and Liu2017; Buatois et al., Reference Buatois, Almond, Mángano, Jensen and Germs2018; Turk et al., Reference Turk, Maloney, Laflamme and Darroch2022), while other studies have shown that the impacts of these behaviors potentially reach Cambrian levels millions of years before the E–C boundary (e.g., Cribb et al., Reference Cribb, Kenchington, Koester, Gibson, Boag, Racicot, Mocke, Laflamme and Darroch2019). When set alongside the apparent low genus diversity of Nama-aged Ediacara biota (Laflamme et al., Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013; Darroch et al., Reference Darroch, Sperling, Boag, Racicot, Mason, Morgan, Tweedt, Myrow, Erwin and Laflamme2015; Reference Darroch, Smith, Laflamme and Erwin2018a,Reference Darroch, Laflamme and Wagnerb), this would seem to lend strong support for the ‘biotic replacement’ model – at least over the White Sea-Nama transition (i.e., the first extinction pulse). However, there remains uncertainty surrounding the extent to which these ecosystem engineering impacts affected soft-bodied Ediacara biota, and thus to what extent there is a predictable pattern of extinction selectivity (Darroch et al., Reference Darroch, Cribb, Buatois, Germs, Kenchington, Smith, Mocke, O’Neil, Schiffbauer, Maloney, Racicot, Turk, Gibson, Almond, Koester, Boag, Tweedt and Laflamme2021).

This point dovetails with the latter, centered on ecological interactions between the Ediacara biota and metazoan fauna. The growing recognition that bed-penetrative bioturbation is much more pervasive and extends further back into the Ediacaran than previously thought (e.g., Jensen et al., Reference Jensen, Saylor, Gehling and Germs2000; Mángano and Buatois, Reference Mángano and Buatois2020; Nelson et al., Reference Nelson, Ramezani, Almond, Darroch, Taylor, Brenner, Furey, Turner and Smith2022; Turk et al., Reference Turk, Maloney, Laflamme and Darroch2022), could suggest that the removal of microbial matgrounds might have been a plausible source of ecological stress and reduction of taphonomically favorable depositional settings. However, not only is there little evidence that matgrounds disappeared over the course of the E–C transition (Buatois et al., Reference Buatois, Narbonne, Mángano, Carmona and Myrow2014), but in addition, frondose Ediacara biota – potentially the group(s) most reliant on matgrounds – are among the taxa that persist right up until the base of the Cambrian. Other hypotheses surrounding the ecosystem engineering impacts of bioturbation also seem to be inconsistent with patterns of extinction and survival over the White Sea-Nama transition, although are hampered by an incomplete understanding of how many groups of Ediacara biota functioned (Darroch et al., Reference Darroch, Cribb, Buatois, Germs, Kenchington, Smith, Mocke, O’Neil, Schiffbauer, Maloney, Racicot, Turk, Gibson, Almond, Koester, Boag, Tweedt and Laflamme2021).

Finally, some work has been done in the area of Question 4 (‘What role did the end-Ediacaran extinction play in the Cambrian Explosion?’), although principally in terms of framing biotic patterns over the E–C transition as an artifact of evolutionary patterns and stem- vs., crown-group diversity dynamics (i.e., suggesting that Ediacaran extinction and Cambrian explosion are different facets of the same process) – see discussion in the above section: ‘Do these intervals of biotic turnover represent extinctions at all?’.

Key questions in 2023

Arguably, one of the most important questions surrounding Ediacaran extinction events is centered on establishing the magnitude of taxonomic loss over the E–C transition (i.e., to what extent these bioevents represent a ‘mass’ extinction) – something that will require better biostratigraphic correlations between fossil localities, and more sophisticated methods for extinction rate modeling. However, we argue that there are a number of other guiding questions which will help build more broad understanding for how pulses of extinction may be part of the sustained interval of biotic innovation that helped sculpt the more modern-looking Phanerozoic biosphere. These are summarized below:

Were extinction pulses slow or rapid?

A crucial question that may eventually help distinguish between ‘catastrophe’ and ‘biotic replacement’ is whether late Ediacaran turnover pulses were rapid, or more protracted. Recent work on the Phanerozoic ‘Big 5’ mass extinctions has shown that the majority of these events were geologically rapid, with community collapse in those events unequivocally driven by environmental perturbation (i.e., the Permian–Triassic and Triassic-Jurassic) occurring on the order of 105 years (Burgess et al., Reference Burgess, Bowring and Shen2014; Erwin, Reference Erwin2014). In a ‘catastrophe’ scenario, therefore, we might expect late Ediacaran turnover pulses to be similarly fast. ‘Biotic replacement’, in contrast, would intuitively be a slower process, arguably operating over a range of evolutionary and ecological timescales as new organisms/behaviors evolve, become successful, disperse, and finally reach widespread ecological significance. Crucially, this model then predicts lengthy stratigraphic overlap between metazoans and Ediacara biota, with diversity decline beginning either once a key behavior emerges, or a threshold in ecosystem engineering intensity is reached. This point, for instance, stands in stark contrast to claims by Wood et al. (Reference Wood, Liu, Bowyer, Wilby, Dunn, Kenchington, Hoyal-Cuthill, Mitchell and Penny2019) that stratigraphic overlap comprises evidence against the ‘biotic replacement’ model. In other words, the lengthy co-existence of Ediacara biota and new ecosystem engineering behaviors is something predicted by ‘biotic replacement’, rather than a criterion for rejecting it. One immediate difficulty with answering this question is the paucity of fossiliferous sections that both span assemblage boundaries and possess sufficient age control. However, recent work on the Dengying Formation in South China has uncovered dickinsoniomorph fossils from near the base of the Shibantan Member (551–543 Ma) below the first occurrence of Cloudina (Xiao et al., Reference Xiao, Chen, Pang, Zhou and Yuan2020; Wang et al., Reference Wang, Chen, Pang, Chuanming, Xiao, Wan and Yuan2021). The Shibantan Member thus likely spans the transition between the White Sea and Nama assemblages (Wang et al., Reference Wang, Chen, Pang, Chuanming, Xiao, Wan and Yuan2021), and may offer an opportunity to constrain the timing of a putative first extinction pulse. Recent work from the Nagoryany Formation in Moldova (Francovschi et al., Reference Francovschi, Gradinaru, Li, Shumlyanskyy and Ciobotaru2021) suggests that these strata preserve a similar interval, and so may offer another opportunity to study this transition in more detail.

What were patterns of extinction selectivity and survivorship over the two turnover pulses?

Extinction events are characterized by ‘victims’ and ‘survivors’, the specific identities of which can offer vital clues as to the source(s) of ecological stress and thus help identify the proximal drivers of extinction. To provide a classic example, Knoll et al. (Reference Knoll, Bambach, Canfield and Grotzinger1996) showed over the Permian–Triassic mass extinction that heavily calcified invertebrates with low metabolic intensities suffered considerably higher extinction intensities than other groups; this pattern suggested that hypercapnia (linked to elevated CO2) was a likely culprit – an inference that pre-empted subsequent work establishing LIP volcanism and ocean acidification as overarching extinction drivers. Thus far, relatively little work has gone into analyzing patterns of selectivity over pulses of E–C extinction, beyond several authors noting that sessile, frondose, and semi-infaunal groups of Ediacara biota overwhelmingly survived the first pulse of extinction at the expense of groups that were mobile and/or surficial. Darroch et al. (Reference Darroch, Cribb, Buatois, Germs, Kenchington, Smith, Mocke, O’Neil, Schiffbauer, Maloney, Racicot, Turk, Gibson, Almond, Koester, Boag, Tweedt and Laflamme2021) noted that sources of ecological stress associated with the specific impacts of bioturbation were hard to ally with the observed extinction selectivity patterns, although there remain big knowledge gaps surrounding what the ecosystem engineering impacts of early metazoans actually were (see question 4 below). In one of the few studies to directly address this question of selectivity, Evans et al. (Reference Evans, Tu, Rizzo, Suprenant, Boan, McCandless, Marshall, Xiao and Droser2022) have noted that the overwhelming survivors of the first extinction pulse – the rangeomorphs and erniettomorphs – are characterized by high surface-area to volume (‘SA:V’) ratios (Laflamme et al., Reference Laflamme, Xiao and Kowalewski2009). With the assumption that high SA:V ratios represent an adaptation (or advantage) to surviving in low-oxygen conditions, then the observed pattern of survivorship would be consistent with extinction driven by fluctuations in global redox conditions (see also Evans et al., Reference Evans, Diamond, Droser and Lyons2018). We note that there are ambiguities surrounding to what extent the erniettomorphs had high surface areas exposed to the water column in life – most seem to have lived at least partially buried in the sediment (Ivantsov et al., Reference Ivantsov, Narbonne, Trusler, Greentree and Vickers-Rich2016; Gibson et al., Reference Gibson, Rahman, Maloney, Racicot, Mocke, Laflamme and Darroch2019; Darroch et al., Reference Darroch, Gibson, Syversen, Rahman, Racicot, Dunn, Gutarra-Diaz, Schindler, Wehrmann and Laflamme2022) – and whether their body plans were evolved for gas exchange as opposed to feeding (see Laflamme et al., Reference Laflamme, Darroch, Tweedt, Peterson and Erwin2013). Despite these caveats, Evans et al. (Reference Evans, Tu, Rizzo, Suprenant, Boan, McCandless, Marshall, Xiao and Droser2022) demonstrate that focusing on selectivity is a powerful means for hypothesis testing, and a more complete knowledge of the physiological, paleoenvironmental, and paleoecological characteristics of ‘victims’ and ‘survivors’ over this boundary could offer vital hints as to what may have driven the first extinction pulse.

The second extinction pulse is similarly enigmatic. Although the vast majority of Ediacara biota disappear at, or shortly beneath, the base of the Cambrian worldwide, there are tantalizing hints that some taxa may persist for short intervals into the lower Cambrian (although we note the difficulties with defining the base of the Cambrian in some of these key sections – e.g., Nelson et al., Reference Nelson, Ramezani, Almond, Darroch, Taylor, Brenner, Furey, Turner and Smith2022). However, there have been no convincing Ediacara biota reported from younger sediments, suggesting that survivors may be ‘dead clades walking’ – a characteristic of several Phanerozoic extinctions whereby some species persist in low numbers through an extinction only to die out early in the recovery interval (Jablonski, Reference Jablonski2002; Hull et al., Reference Hull, Darroch and Erwin2015). Among putative survivors, three examples stand out as being particularly convincing and/or in need of closer analysis: Jensen et al. (Reference Jensen, Gehling and Droser1998) figure apparent erniettomorph Ediacara biota from the Uratanna Formation in South Australia, stratigraphically above the FAD of T. pedum. In the southwestern United States, Hagadorn and Waggoner (Reference Hagadorn and Waggoner2000) report forms similar to the erniettomorph taxon Swartpuntia from two separate localities above the FAD of T. pedum, and from strata that also preserve trilobites and archaeocyaths. Most recently, Nelson et al. (Reference Nelson, Ramezani, Almond, Darroch, Taylor, Brenner, Furey, Turner and Smith2022) report an erniettomorph from the Nomtsas Formation in the Neint Nababeep Plateau in South Africa. Although they interpreted this occurrence as Ediacaran, given that the Nomtsas Formation is traditionally interpreted as Cambrian (e.g., Linnemann et al., Reference Linnemann, Ovtcharova, Schaltegger, Gartner, Hautmann, Geyer, Vickers-Rich, Rich, Plessen, Hofmann, Zieger, Krause, Kriesfield and Smith2019), this may represent yet another survivor. Although these data are sparse (and have yet to be fully investigated beyond an initial description), a preliminary analysis of selectivity across this second extinction pulse may suggest that frondose erniettomorphs were survivors, while rangeomorph Ediacara biota were permanent casualties (although see Hoyal Cuthill, Reference Hoyal Cuthill2022 for an alternative view of rangeomorph extinction). If it is shown that erniettomorphs had significantly different paleobiologies or -ecologies than these other groups, then this may help establish extinction drivers over the E–C boundary itself.

Are Ediacaran–Cambrian carbon isotope excursions recording global environmental perturbations?

Carbon isotope excursions in the late Ediacaran and earliest Cambrian are, largely, interpreted as recording perturbations to the global dissolved inorganic carbon (DIC) reservoir. As such, δ13C chemostratigraphy from carbonate successions has been used to correlate between sites regionally and globally, to construct age models, and to calibrate biostratigraphic changes across the E–C transition, particularly in the absence of radioisotopically constrained sections (e.g., Maloof et al., Reference Maloof, Porter, Moore, Dudas, Bowring, Higgins, Fike and Eddy2010; Bowyer et al., Reference Bowyer, Zhuravlev, Wood, Shields, Zhou, Curtis, Poulton, Condon, Yang and Zhu2022). Although this approach has been applied widely to E–C studies, a number of geochemical studies of modern and ancient carbonate platforms have demonstrated the pitfalls of indiscriminate use of carbon isotope chemostratigraphy. Studies of modern carbonate platforms have shown that mineralogy and early marine diagenesis in shallow marine environments can result in variable δ13C records that are recording sediment- and fluid-buffered diagenetic precipitates that can be decoupled from global DIC (Swart, Reference Swart2008; Oehlert and Swart, Reference Oehlert and Swart2014; Higgins et al., Reference Higgins, Blattler, Lundstrom, Santiago-Ramos, Akhtar, Ahm, Bialik, Holmden, Bradbury, Murray and Swart2018). Building upon these modern studies, chemostratigraphic data from Neoproterozoic carbonate platforms have demonstrated lateral δ13C variability across shelf to slope transects and, using the interpretive framework established in studies of modern carbonate platforms, have interpreted some of this variability as the result of early diagenesis (Ahm et al., Reference Ahm, Maloof, Macdonald, Hoffman, Bjerrum, Bold, Rose, Strauss and Higgins2019; Hoffman and Lamothe, Reference Hoffman and Lamothe2019). Other studies of partially dolomitized Neoproterozoic platforms have demonstrated that dolomitizing fluids have the potential to alter δ13C values by up to 10‰ (Bold et al., Reference Bold, Ahm, Schrag, Higgins, Jamsram and Macdonald2020; Nelson et al., Reference Nelson, Ahm, Macdonald, Higgins and Smith2021). Finally, some detailed stratigraphic investigations have demonstrated a facies dependence on the character and/or preservation of carbon isotope excursions (e.g., Lu et al., Reference Lu, Zhu, Zhang, Shields-Zhou, Li, Zhao, Zhao and Zhao2013; Busch et al., Reference Busch, Hodgin, Ahm, Husson, Macdonald, Bergmann, Higgins and Strauss2022). Collectively, these studies highlight the need for careful assessment of the diagenetic histories of individual E–C carbonate platforms before records from individual sites can be interpreted within a global framework. With the recognition of more E–C carbon isotope excursions, some of which are only convincingly documented in a single region (see Figure 1), this regional scale assessment of diagenetic history is particularly important.

A parallel challenge for the community will be linking individual negative carbon isotope excursions that are established as global, to viable environmental perturbations that could result in an influx of light carbon. This challenge is not a new one. The extreme fluctuations in δ13C that characterize much of the Neoproterozoic have long been difficult to interpret in a mass-balance framework because of the dramatic changes in oxidants that are implied. Despite these long-standing challenges, rifting and rift-related volcanism of broadly E–C age occurred around the margins of Laurentia (Figure 3) and, as with some of the volcanic episodes associated with Phanerozoic mass extinctions, have recently been proposed as a possible “trigger” for a cascade of E–C environmental, ecological, and biotic effects (Hodgin et al., Reference Hodgin, Nelson, Wall, Barron-Diaz, Webb, Schmitz, Fike, Hagadorn and Smith2021; Smith et al., Reference Smith, Nelson, O’Connell, Eyster and Lonsdale2022). Similar to the historical trajectory of the study of Phanerozoic mass extinctions (e.g., Newell, Reference Newell1967), the first step in testing the idea that E–C rift-related volcanism caused a perturbation(s) to geochemical cycles and a biotic crisis is demonstrating temporal coincidence among them. After temporal coincidence is established, the focus can turn to studying the geochemical response and the expected biotic selection across the E–C transition.

What were the ecosystem engineering impacts of early animals?

The question as to how Ediacara biota and metazoans were interacting (discussed above) has yet to be fully answered, however, there are arguably more fundamental questions surrounding what the ecosystem engineering impacts of early animals actually were, and whether they were capable of driving large scale environmental change. While Cribb et al. (Reference Cribb, Kenchington, Koester, Gibson, Boag, Racicot, Mocke, Laflamme and Darroch2019) focused on trace fossils and quantified their effects as indices of ‘ecosystem engineering impact’ (the ‘EEIs’ of Herringshaw et al., Reference Herringshaw, Callow and McIlroy2017), these may overestimate downstream effects, and so their use has been criticized (Minter et al., Reference Minter, Buatois, Mangano, Davies, Gibling, Macnaughton and Labandeira2017). Other workers have highlighted the importance of biomixing vs. bioirrigation. For clarification, biomixing has relatively little ecosystem engineering impact (especially at shallow depths), whereas bioirrigation is a more powerful driver of environmental change – leading to deepening redox gradients, altered distribution of redox-sensitive elements, and increased availability of organic matter (e.g., Kristensen et al., Reference Kristensen, Penha-Lopes, Delefosse, Valdemarsen, Quintana and Banta2012; Tarhan et al., Reference Tarhan, Droser, Planavsky and Johnston2015, Reference Tarhan, Zhao and Planavsky2021; Darroch et al., Reference Darroch, Cribb, Buatois, Germs, Kenchington, Smith, Mocke, O’Neil, Schiffbauer, Maloney, Racicot, Turk, Gibson, Almond, Koester, Boag, Tweedt and Laflamme2021). In this regard, the presence of treptichnid burrows in the late Ediacaran (Cribb et al., Reference Cribb, Kenchington, Koester, Gibson, Boag, Racicot, Mocke, Laflamme and Darroch2019; Jensen et al., Reference Jensen, Saylor, Gehling and Germs2000; Darroch et al., Reference Darroch, Cribb, Buatois, Germs, Kenchington, Smith, Mocke, O’Neil, Schiffbauer, Maloney, Racicot, Turk, Gibson, Almond, Koester, Boag, Tweedt and Laflamme2021) is crucial; treptichnids record the first substantial bioirrigative behaviors to appear in the trace fossil record (likely between 542.65–539.63 Ma; see age model in Nelson et al., Reference Nelson, Ramezani, Almond, Darroch, Taylor, Brenner, Furey, Turner and Smith2022) and rapidly increase in size, complexity, and intensity through the latest Ediacaran and into the Cambrian. However, the actual downstream effects of this style of burrowing have not been measured, and so the significance of Ediacaran treptichnid-like behaviors is unknown. What is needed is a combination of in vivo ichnological experiments (encompassing a wide variety of different tracemakers) with integrated geochemical models; with these two approaches we might reasonably hope to understand what the significance of these behavioral innovations may have been (see, for example, Cribb et al., Reference Cribb, van de Velde, Berelson, Bottjer and Corsetti2023).

Lastly, in addition to the record of bioturbation, there is a wealth of other ecosystem engineering impacts that appear at approximately the same time and should be more broadly considered (e.g., Erwin and Tweedt, Reference Erwin and Tweedt2012). Recent studies using fluid dynamics modeling have noted a paleoecological shift in the prevalence and character of inferred suspension feeders from the White Sea into the Nama intervals (Gibson et al., Reference Gibson, Rahman, Maloney, Racicot, Mocke, Laflamme and Darroch2019;Cracknell et al., Reference Cracknell, Ankor, García-Bellido, Gehling, Darroch and Rahman2021; Darroch et al., Reference Darroch, Gibson, Syversen, Rahman, Racicot, Dunn, Gutarra-Diaz, Schindler, Wehrmann and Laflamme2022), which could have altered resource flows, and the distribution of habitable ecospace. Alternatively, it is also becoming apparent that passive predation is a mode of life that perhaps emerged as early as the Avalon assemblage (Liu et al., Reference Liu, Matthews, Menon, McIlroy and Brasier2014; Dunn et al., Reference Dunn, Kenchington, Parry, Clark, Kendall and Wilby2022), but then may have expanded dramatically in the Nama (Bengston and Zhao, Reference Bengston and Zhao1992; Darroch et al., Reference Darroch, Boag, Racicot, Tweedt, Mason, Erwin and Laflamme2016; Schiffbauer et al., Reference Schiffbauer, Huntley, O’Neil, Darroch, Laflamme and Cai2016; Leme et al., Reference Leme, Van Iten and Simoes2022; Turk et al., Reference Turk, Maloney, Laflamme and Darroch2022). Although not strictly an ecosystem engineering impact, the rise of predation was likely a source of ecological antagonistic stress that could have gradually marginalized the Ediacara biota, particularly if any of these groups possessed a mobile larval or dispersal stage early in development (Darroch et al., Reference Darroch, Boag, Racicot, Tweedt, Mason, Erwin and Laflamme2016). A prediction of this might be that Ediacaran communities from the Avalon through Nama assemblages show a noticeable shift from more neutral to niche-dominated processes in ecosystem dynamics – a facet of community paleoecology that can be readily preserved in the spatial distributions of fossils on bedding planes (see Mitchell et al., Reference Mitchell, Harris, Kenchington, Vixseboxse, Roberts, Clark, Dennis, Liu and Wilby2019, Reference Mitchell, Evans, Chen and Xiao2022).

Bonus question: What were the ecosystem engineering impacts of Ediacara biota?

Finally, much discussion surrounding the ‘biotic replacement’ model has focused on the impact of metazoans and the emerging Cambrian-style fauna as ecosystem engineers. However, this conceptual approach ignores to what extent the evolving Ediacara biota may have been engaged in forms of ecosystem engineering themselves, and thus may be reinforcing a false narrative surrounding the character and drivers of turnover. Although none of the Ediacara biota are thought to have disrupted the sediment–water interface to the extent that Cambrian-style metazoans did, many might have engaged in other forms of ecosystem engineering. For example, as mentioned above, several groups of Ediacara biota were likely suspension feeders (Rahman et al., Reference Rahman, Darroch, Racicot and Laflamme2015; Gibson et al., Reference Gibson, Rahman, Maloney, Racicot, Mocke, Laflamme and Darroch2019) and thus may have helped fuel the Cambrian explosion through forging energetic links between pelagic and benthic realms (Cracknell et al., Reference Cracknell, Ankor, García-Bellido, Gehling, Darroch and Rahman2021; Darroch et al., Reference Darroch, Gibson, Syversen, Rahman, Racicot, Dunn, Gutarra-Diaz, Schindler, Wehrmann and Laflamme2022), as well as help oxygenate the water column (Erwin and Tweedt, Reference Erwin and Tweedt2012). In addition, other groups of sessile – predominantly frondose – Ediacara biota apparently formed dense ‘meadows’ in both shallow and deep-water settings (see, e.g., Clapham et al., Reference Clapham, Narbonne and Gehling2003; Droser and Gehling, Reference Droser and Gehling2008) which would have baffled currents, altered resource flows, and created a diversity of benthic niches and hydrodynamic refugia that could have been exploited by other taxa (analogous to modern seagrass meadows – see, e.g., Gartner et al., Reference Gartner, Tuya, Lavery and McMahon2013). Lastly, Budd and Jensen (Reference Budd and Jensen2017) have suggested that, following death and decay, patches of sessile Ediacara biota may have served as rich sources of organic matter – similar to whale falls – that in turn would have provided a selective pressure towards motility and the development of deposit-feeding strategies. Although some of these models are more plausible than others, they collectively emphasize that little is currently known about the ecosystem engineering impacts of the Ediacara biota. More broadly, these ideas illustrate that a key facet of understanding pulses of late Ediacaran extinction – both in terms of testing between hypothesized drivers and analyzing extinction selectivity – is understanding what different groups of Ediacara biota were actually doing within their ecosystems.

Summary

Pulses of extinction in the late Ediacaran remain among the most enigmatic events in the history of life – occurring at a crucial interval during the Neoproterozoic rise of animals – and thus potentially influencing trends in early animal evolution as well as the character of the emerging, more modern-functioning and animal-dominated marine biosphere. The extent to which these events represent ‘mass extinctions’ – that is, the rapid disappearance of >70% of marine genera, rather than less severe and more protracted turnover pulses, is a question that doubtless requires a more focused analysis. Inferring the drivers of these putative extinction pulses is also a question that is fraught with difficulties stemming from biostratigraphic correlation and the interpretation of enigmatic geochemical signals. Despite this, the last 5 years of research has produced a wealth of new geological, paleontological, geochemical, and geochronological data that are slowly bringing this interval into focus. These new data have answered several of the questions posed by Darroch et al. (Reference Darroch, Smith, Laflamme and Erwin2018a), but have left others unanswered. Moreover, this work has led to new questions which promise to not only help unravel this critical interval in Earth’s history, but also contribute to our knowledge of extinction events more generally. Addressing the questions listed above will: (1) help link sources of environmental change in the late Ediacaran with drivers of ecological stress; and (2) explain patterns of extinction and survivorship across extinction pulses. Taken together, this new information will allow for a coherent picture of late Ediacaran extinction and help determine the role it may have played in the Cambrian explosion.

Finally, the last 5 years of research allow us to propose a revised hypothesis for drivers of E–C biotic turnover, that can be tested and refined with further discoveries. Modeled after Knoll and Carroll (Reference Knoll and Carroll1999), the various tenets of this hypothesis build off a combination of growing consensus and current ideas, specifically: (1) the majority of ‘Ediacara biota’ are stem-group members of extant animal phyla, with crown group members of these same clades emerging as early as the Avalon assemblage (e.g., Dunn et al., 2018; Dunn et al., Reference Dunn, Kenchington, Parry, Clark, Kendall and Wilby2022); (2) in the absence of temporal correlation between isotope excursions and diversity loss (Rooney et al., Reference Rooney, Cantine, Bergmann, Gomez-Perez, Al Baloushi, Boag, Busch, Sperling and Strauss2020), a first pulse of extinction is recognized at the White Sea-Nama transition driven by competition, biotic interactions, and widespread geobiological change stemming from the diversification of crown-group metazoan clades (i.e., ‘biotic replacement’; see Darroch et al., Reference Darroch, Sperling, Boag, Racicot, Mason, Morgan, Tweedt, Myrow, Erwin and Laflamme2015; Schiffbauer et al., Reference Schiffbauer, Huntley, O’Neil, Darroch, Laflamme and Cai2016); and (3) a second pulse of extinction at the E–C boundary, potentially driven by widespread environmental perturbation (i.e., ‘catastrophe’) following extensive rift volcanism around the southern and southwestern margins of Laurentia (Hodgin et al., Reference Hodgin, Nelson, Wall, Barron-Diaz, Webb, Schmitz, Fike, Hagadorn and Smith2021; Smith et al., Reference Smith, Nelson, O’Connell, Eyster and Lonsdale2022) (Figure 4). This model thus hypothesizes a two-pulsed extinction of the Ediacara biota – with a first pulse driven by ‘biotic replacement’ and a second pulse driven by ‘catastrophe’ (i.e., environmental perturbation), combining recent evidence from paleontology, geochronology, and geochemistry. This model is obviously highly speculative and sensitive to the questions outlined above, but does frame testable hypotheses that can, and will, be targeted in the next decade of work. Testing this model will allow us to fold the biotic turnover events occurring over the E–C transition into broader discussions surrounding the tempo, mode, and drivers of mass extinction events, and invite comparisons with the ‘Big 5’ mass extinctions of the Phanerozoic.

Figure 4. A hypothetical model for drivers of the E–C transition, modeled after Knoll and Carroll (Reference Knoll and Carroll1999); their figure 5). ‘AV’ = Avalon; ‘WS’ = White Sea; ‘NAM’ = Nama. This model interprets the majority of ‘Ediacara biota’ as stem-group metazoans, with a first pulse of extinction at the White Sea-Nama transition driven by the diversification – with associated downstream geobiological impacts – of crown-group metazoan clades (i.e., ‘biotic replacement’). A second pulse of extinction follows at the E–C boundary, driven by widespread environmental perturbation (i.e., ‘catastrophe’) following extensive rift volcanism. We note that this figure is strictly hypothetical, and the stem – and crown-groups depicted here do not intentionally correspond to specific biological groups shown in Figure 1.

Open peer review

To view the open peer review materials for this article, please visit http://doi.org/10.1017/ext.2023.12.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/ext.2023.12.

Acknowledgements

S.A.F.D. and M.C. are grateful to the Ecological and Evolutionary Effects of Extinction and Ecosystem Engineers (‘E6’) group for helpful discussions and feedback. We thank our reviewers and handling editor for thoughtful comments and discussion surrounding E–C extinction rates, and which considerably improved an earlier version of this manuscript.

Author contribution

Conceptualization: all authors.; Formal analysis (principally Figure 1 and calculation of extinction intensities): M.C.; Funding acquisition: S.A.F.D., E.F.S., M.L. and J.D.S.; Writing—original draft: S.A.F.D.; Writing—review & editing: all authors.

Financial support

S.A.F.D. was supported by a National Science Foundation RCN grant (NSF-DEB 2051255), and joint funding from the National Science Foundation (NSF EAR-2007928) and Natural Environment Research Council (NE/V010859/2). S.A.F.D. also acknowledges generous support from the Alexander von Humboldt Foundation, which is sponsored by the Federal Ministry for Education and Research in Germany. M.L. was supported by an NSERC Discovery Grant (RGPIN 435402). E.F.S. acknowledges support from the National Science Foundation (NSF EAR-2144836 and NSF EAR-2021064), the Sloan Research Fellowship (#FG-2021-2116,049), and the Johns Hopkins Catalyst Award. This is E6 (Ecological and Evolutionary Effects of Extinction and Ecosystem Engineers RCN) publication #2. J.D.S. acknowledges support from the National Science Foundation (NSF EAR CAREER-1652351).

Competing interest

The authors declare none.

References

Ahm, A-SC, Maloof, AC, Macdonald, FA, Hoffman, PF, Bjerrum, CJ, Bold, U, Rose, CV, Strauss, JV and Higgins, JA (2019) An early diagenetic deglacial origin for basal Ediacaran “cap dolostones”. Earth and Planetary Science Letters 506, 292307.CrossRefGoogle Scholar
Algeo, TJ and Scheckler, SE (1998) Terrestrial-marine teleconnections in the Devonian: Links between the evolution of land plants, weathering processes, and marine anoxic events. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences 353, 113130.CrossRefGoogle Scholar
Alroy, J (2008) Dynamics of origination and extinction in the marine fossil record. Proceedings of the National Academy of Sciences 105(supplement 1), 1153611542.CrossRefGoogle ScholarPubMed
Alroy, J (2014) Accurate and precise estimates of origination and extinction rates. Paleobiology 40, 374397.CrossRefGoogle Scholar
Amthor, JE, Grotzinger, JP, Shröder, S, Bowring, SA, Ramezani, J, Martin, MW and Matter, A (2003) Extinction of Cloudina and Namacalathus at the Precambrian–Cambrian boundary in Oman. Geology 31, 431434.2.0.CO;2>CrossRefGoogle Scholar
Bengston, S and Zhao, Y (1992) Predatorial borings in late Precambrian mineralized exoskeletons. Science 257, 367369.Google Scholar
Boag, TH, Darroch, SAF and Laflamme, M (2016) Ediacaran distributions in space and time: Testing assemblage concepts of earliest macroscopic body fossils. Paleobiology 42, 574594.CrossRefGoogle Scholar
Boddy, CE, Mitchell, EG, Merdith, A and Liu, AG (2022) Palaeolatitudinal distribution of the Ediacaran macrobiota. Journal of the Geological Society of London 179, jgs2021–030.CrossRefGoogle Scholar
Bold, U, Ahm, A-SC, Schrag, DP, Higgins, JA, Jamsram, E and Macdonald, FA (2020) Effect of dolomitization on isotopic records from Neoproterozoic carbonates in southwestern Mongolia. Precambrian Research 350, 105902.CrossRefGoogle Scholar
Bowring, SA, Grotzinger, JP, Condon, DJ, Ramezani, J, Newall, MJ and Allen, PA (2007) Geochronologic constraints on the chronostratigraphic framework of the Neoproterozoic Huqf Supergroup, Sultanate of Oman. American Journal of Science 307, 10971145.CrossRefGoogle Scholar
Bowyer, FT, Zhuravlev, AY, Wood, R, Shields, GA, Zhou, Y, Curtis, A, Poulton, SW, Condon, DJ, Yang, C and Zhu, M (2022) Calibrating the temporal and spatial dynamics of the Ediacaran-Cambrian radiation of animals. Earth Science Reviews 225, 103913.CrossRefGoogle Scholar
Brasier, MD (1992) Background to the Cambrian explosion. Journal of the Geological Society 149, 585587.CrossRefGoogle Scholar
Brasier, M, Cowie, J and Taylor, M (1994) Decision on the Precambrian-Cambrian boundary stratotype. Episodes Journal of International Geoscience 17, 38.Google Scholar
Buatois, LA, Almond, J, Mángano, MG, Jensen, S and Germs, GJB (2018) Sediment disturbance by Ediacaran bulldozers and the roots of the Cambrian explosion. Scientific Reports 8, 4514.CrossRefGoogle ScholarPubMed
Buatois, LA, Narbonne, GM, Mángano, MG, Carmona, NB and Myrow, P (2014) Relict ecosystems at the dawn of the Phanerozoic revolution. Nature Communications 5, 35443549.CrossRefGoogle Scholar
Budd, GE and Jensen, S (2017) The origin of the animals and a ‘savannah’ hypothesis for early bilaterian evolution. Biological Reviews 92, 446473.CrossRefGoogle Scholar
Budd, GE and Mann, RP (2020) The dynamics of stem and crown groups. Science Advances 6, eaaz1626.CrossRefGoogle ScholarPubMed
Burgess, SD, Bowring, S and Shen, S (2014) High-precision timeline for Earth’s most severe extinction. Proceedings of the National Academy of Sciences 111, 33163321.CrossRefGoogle ScholarPubMed
Busch, JF, Hodgin, EB, Ahm, A-SC, Husson, JM, Macdonald, FA, Bergmann, KD, Higgins, JA and Strauss, JV (2022) Global and local drivers of the Ediacaran Shuram carbon isotope excursion. Earth and Planetary Science Letters 579, 117368.CrossRefGoogle Scholar
Cai, Y, Xiao, S, Li, G and Hua, H (2019) Diverse biomineralizing animals in the terminal Ediacaran period herald the Cambrian explosion. Geology 47, 380384.CrossRefGoogle Scholar
Cañadas, F, Papineau, D, Leng, ML and Li, C (2022) Extensive primary production promoted the recovery of the Ediacaran Shuram excursion. Nature Communications 13, 148.CrossRefGoogle ScholarPubMed
Canfield, De, Knoll, AH, Poulton, SW, Narbonne, GM and Dunning, GAR (2020) Carbon isotopes in clastic rocks and the Neoproterozoic carbon cycle. American Journal of Science 320, 97124.CrossRefGoogle Scholar
Chen, Z, Zhou, C, Yuan, X and Xiao, S (2019) Death march of a segmented and trilobate bilaterian elucidates early animal evolution. Nature 573, 412415.CrossRefGoogle ScholarPubMed
Clapham, ME, Narbonne, GM and Gehling, JG (2003) Paleoecology of the oldest-known animal communities: Ediacaran assemblages at mistaken point, Newfoundland. Paleobiology 29, 527544.2.0.CO;2>CrossRefGoogle Scholar
Clarkson, MO, Kasemann, SA, Wood, RA, Lenton, TM, Daines, SJ, Richoz, S, Ohnemueller, F, Meixner, A, Poulton, SW and Tipper, ET (2015) Ocean acidification and the Permo-Triassic mass extinction. Science 348, 229232.CrossRefGoogle ScholarPubMed
Cracknell, K, Ankor, MJ, García-Bellido, DC, Gehling, JG, Darroch, SAF and Rahman, IA (2021) Pentaradial eukaryote suggests expansion of suspension feeding in White Sea-aged Ediacaran communities. Scientific Reports 11, 4121.CrossRefGoogle ScholarPubMed
Cribb, AT, Kenchington, CG, Koester, B, Gibson, BM, Boag, TH, Racicot, RA, Mocke, H, Laflamme, M and Darroch, SAF (2019) Increase in metazoan ecosystem engineering prior to the Ediacaran-Cambrian boundary in the Nama group, Namibia. Royal Society Open Science 6, 190548.CrossRefGoogle Scholar
Cribb, AT, van de Velde, J, Berelson, WM, Bottjer, DJ and Corsetti, FA (2023) Ediacaran-Cambrian bioturbation did not extensively oxygenate sediments in shallow-marine ecosystems. Geobiology 119. https://onlinelibrary.wiley.com/toc/14724669/0/0Google Scholar
Darroch, SAF, Boag, T, Racicot, RA, Tweedt, S, Mason, SJ, Erwin, DH and Laflamme, M (2016) A mixed Ediacaran-metazoan assemblage from the Zaris sub-basin, Namibia. Palaeogeography, Palaeoclimatology, Palaeoecology 459, 198208.CrossRefGoogle Scholar
Darroch, SAF, Cribb, AT, Buatois, LA, Germs, GJB, Kenchington, CG, Smith, EF, Mocke, H, O’Neil, GR, Schiffbauer, JD, Maloney, KM, Racicot, RA, Turk, KA, Gibson, BM, Almond, J, Koester, B, Boag, TM, Tweedt, SM and Laflamme, M (2021) The trace fossil record of the Nama group, Namibia: Exploring the terminal Ediacaran roots of the Cambrian explosion. Earth Science Reviews 212, 103435.CrossRefGoogle Scholar
Darroch, SAF, Gibson, BM, Syversen, M, Rahman, IA, Racicot, RA, Dunn, FS, Gutarra-Diaz, S, Schindler, E, Wehrmann, A and Laflamme, M (2022) The life and times of Pteridinium simplex. Paleobiology 48, 527556.CrossRefGoogle Scholar
Darroch, SAF, Laflamme, M and Wagner, PJ (2018b) High ecological complexity in benthic Ediacaran ecosystems. Nature Ecology and Evolution 2, 15411547.CrossRefGoogle Scholar
Darroch, SAF, Smith, EF, Laflamme, M and Erwin, DH (2018a) Ediacaran extinction and Cambrian explosion. Trends in Ecology and Evolution 33, 653663.CrossRefGoogle ScholarPubMed
Darroch, SAF, Sperling, EA, Boag, T, Racicot, RA, Mason, SJ, Morgan, AS, Tweedt, S, Myrow, P, Erwin, DH and Laflamme, M (2015) Biotic replacement and mass extinction of the Ediacara biota. Proceedings of the Royal Society B 282, 20151003.CrossRefGoogle ScholarPubMed
Droser, ML and Gehling, JG (2008) Synchronous aggregate growth in an abundant new Ediacaran tubular organism. Science 319, 16601662.CrossRefGoogle Scholar
Droser, ML and Gehling, JG (2015) The advent of animals: The view from the Ediacaran. Proceedings of the National Academy of Sciences 112, 48654870.CrossRefGoogle ScholarPubMed
Droser, ML, Tarhan, LG and Gehling, JG (2017) The rise of animals in a changing environment: global ecological innovation in the late Ediacaran. Annual Review of Earth and Planetary Sciences 45, 593617.CrossRefGoogle Scholar
Dunn, FS, Kenchington, CG, Parry, LA, Clark, JW, Kendall, RS and Wilby, PR (2022) A crown-group cnidarian from the Ediacaran of Charnwood Forest, UK. Nature Ecology and Evolution 6, 10951104.CrossRefGoogle ScholarPubMed
Dunn, FS, Liu, AG, Grazhdankin, DV, Vixseboxse, P, Flanerry-Sutherland, J, Green, E, Harris, S, Wilby, PR and Donoghue, PCJ (2021) The developmental biology of Charnia and the eumetazoan affinity of the Ediacaran rangeomorphs. Science Advances 7, eabe0291.CrossRefGoogle Scholar
Eden, R, Manica, A and Mitchell, EG (2022) Metacommunity analyses show an increase in ecological specialisation throughout the Ediacaran period. PLoS Biology 20, e3001289.CrossRefGoogle ScholarPubMed
Erwin, DH (2008) Macroevolution of ecosystem engineering, niche construction and diversity. Trends in Ecology and Evolution 23, 304310.CrossRefGoogle ScholarPubMed
Erwin, DH (2014) Temporal acuity and the rate and dynamics of mass extinctions. Proceedings of the National Academy of Sciences 111, 32033204.CrossRefGoogle Scholar
Erwin, DH, Laflamme, M, Tweedt, SM, Sperling, EA, Pisani, D and Peterson, KJ (2011) The Cambrian conundrum: Early divergence and later ecological success in the early history of animals. Science 334, 10911097.CrossRefGoogle ScholarPubMed
Erwin, DH and Tweedt, SM (2012) Ecological drivers of the Ediacaran-Cambrian diversification of Metazoa. Evolutionary Ecology 26, 417433.CrossRefGoogle Scholar
Evans, SD, Diamond, CW, Droser, ML and Lyons, TW (2018) Dynamic oxygen and coupled biological and ecological innovation during the second wave of the Ediacara biota. Emerging Topics in Life Sciences 2, 223233.Google ScholarPubMed
Evans, SD, Tu, C, Rizzo, A, Suprenant, RL, Boan, PC, McCandless, H, Marshall, N, Xiao, S and Droser, ML (2022) Environmental drivers of the first major animal extinction across the Ediacaran White Sea-Nama transition. Proceedings of the National Academy of Sciences 119, e2207475119.CrossRefGoogle ScholarPubMed
Fike, DA, Grotzinger, JP, Pratt, LM and Summons, RE (2006) Oxidation of the Ediacaran ocean. Nature 444, 744747.CrossRefGoogle ScholarPubMed
Foote, M (1999) Morphological diversity in the evolutionary radiation of Paleozoic and post-Paleozoic crinoids. Paleobiology 25, 1115.CrossRefGoogle Scholar
Francovschi, I, Gradinaru, E, Li, H, Shumlyanskyy, L and Ciobotaru, V (2021) U–Pb geochronology and Hf isotope systematics of detrital zircon from the late Ediacaran Kalyus beds (east European platform): Palaeogeographic evolution of southwestern Baltica and constraints on the Ediacaran biota. Precambrian Research 355, 106062.CrossRefGoogle Scholar
Gartner, A, Tuya, F, Lavery, PS and McMahon, K (2013) Habitat preferences of macroinvertebrate fauna among seagrasses with varying structural forms. Journal of Experimental Marine Biology and Ecology 439, 143151.CrossRefGoogle Scholar
Gehling, JG (1991) The case for Ediacaran fossil roots to the metazoan tree. Geological Society of India Memoir 20, 181224.Google Scholar
Gehling, JG (1999) Microbial mats in terminal Proterozoic siliciclastics: Ediacaran death masks. PALAIOS 14, 4057.CrossRefGoogle Scholar
Germs, GJB (1972) New shelly fossils from the Nama group, south West Africa. American Journal of Science 272, 752761.CrossRefGoogle Scholar
Gibson, BM, Rahman, I, Maloney, K, Racicot, R, Mocke, H, Laflamme, M and Darroch, SAF (2019) Gregarious suspension feeding in a modular Ediacaran organism. Science Advances 5, eaaw0260.CrossRefGoogle Scholar
Gibson, BM, Schiffbauer, JD, Wallace, AF and Darroch, SAF (2023) The role of iron in the formation of Ediacaran ‘death masks’. Geobiology 114.Google ScholarPubMed
Glaessner, MF (1984) The Dawn of Animal Life: A Biohistorical Study. New York: Cambridge University Press.Google Scholar
Gold, DA, Runnegar, B, Gehling, JG and Jacobs, DK (2015) Ancestral state reconstruction of ontogeny supports a bilaterian affinity for Dickinsonia. Evolution and Development 17, 315324.CrossRefGoogle ScholarPubMed
Grazhdankin, D (2004) Patterns of distribution in the Ediacaran biotas: Facies versus biogeography and evolution. Paleobiology 30, 203221.2.0.CO;2>CrossRefGoogle Scholar
Grazhdankin, DV (2014) Patterns of evolution of the Ediacaran soft-bodied biota. Journal of Paleontology 88, 269283.CrossRefGoogle Scholar
Grotzinger, JP, Fike, DA and Fischer, WW (2011) Enigmatic origin of the largest-known carbon isotope excursion in Earth’s history. Nature Geoscience 4, 285292.CrossRefGoogle Scholar
Hagadorn, JW and Waggoner, B (2000) Ediacaran fossils from the southwestern Great Basin, United States. Journal of Paleontology 74, 349359.2.0.CO;2>CrossRefGoogle Scholar
Herringshaw, LG, Callow, RT and McIlroy, D (2017) Engineering the Cambrian explosion: The earliest bioturbators as ecosystem engineers. Geological Society of London Special Publications 448, 369382.CrossRefGoogle Scholar
Higgins, JA, Blattler, CL, Lundstrom, EA, Santiago-Ramos, D, Akhtar, A, Ahm, A-SC, Bialik, O, Holmden, C, Bradbury, H, Murray, ST and Swart, P (2018) Mineralogy, early marine diagenesis, and the chemistry of shallow water carbonate sediments. Geochimica et Cosmochimica Acta 220, 512534.CrossRefGoogle Scholar
Hodgin, EB, Nelson, LL, Wall, CJ, Barron-Diaz, AJ, Webb, LC, Schmitz, MD, Fike, DA, Hagadorn, JW and Smith, EF (2021) A link between rift-related volcanism and end-Ediacaran extinction? Integrated chemostratigraphy, biostratigraphy, and U-Pb geochronology from Sonora, Mexico. Geology 49, 115119.CrossRefGoogle Scholar
Hoffman, PF and Lamothe, KG (2019) Seawater-buffered diagenesis, destruction of carbon isotope excursions, and the composition of DIC in Neoproterozoic oceans. Proceedings of the National Academy of Sciences USA 116, 1887418879.CrossRefGoogle ScholarPubMed
Hoyal Cuthill, J (2022) Ediacaran survivors in the Cambrian: suspicions, denials and a smoking gun. Geological Magazine 159, 12101219.CrossRefGoogle Scholar
Hull, PM, Darroch, SAF and Erwin, DH (2015) Rarity in mass extinctions and the future of ecosystems. Nature 528, 345351.CrossRefGoogle ScholarPubMed
Husson, JM, Higgins, JA, Maloof, AC and Schoene, B (2015) Ca and mg isotope constraints on the origin of Earth’s deepest δ13C excursion. Geochimica et Cosmochimica Acta 160, 243266.CrossRefGoogle Scholar
Ireland, TR, Flöttmann, T, Fanning, CM, Gibson, GM and Preiss, WV (1998) Development of the early Paleozoic Pacific margin of Gondwanafrom detrital-zircon ages across the Delamerian orogen. Geology 26, 243246.2.3.CO;2>CrossRefGoogle Scholar
Ivantsov, AY, Narbonne, GM, Trusler, PW, Greentree, C and Vickers-Rich, P (2016) Elucidating Ernietta: new insights from exceptional specimens in the Ediacaran of Namibia. Lethaia 49, 540554.CrossRefGoogle Scholar
Jablonski, D (2002) Survival without recovery after mass extinctions. Proceedings of the National Academy of Sciences USA 99, 81398144.CrossRefGoogle ScholarPubMed
Jensen, S, Gehling, JG and Droser, ML (1998) Ediacara-type fossils in Cambrian sediments. Nature 393, 567569.CrossRefGoogle Scholar
Jensen, S, Saylor, BZ, Gehling, JG and Germs, GJB (2000) Complex trace fossils from the terminal Proterozoic of Namibia. Geology 28, 143146.2.0.CO;2>CrossRefGoogle Scholar
Jones, CG, Lawton, JH and Shachak, M (1994) Organisms as ecosystem engineers. Oikos 69, 373386.CrossRefGoogle Scholar
Kirschvink, JL, Magaritz, M, Ripperdan, RL, Zhuravlev, AY and Rozanov, AY (1991) The Precambrian-Cambrian boundary: Magnetostratigraphy and carbon isotopes resolve correlation problems between Siberia, Morocco, and South China. GSA Today 1, 6991.Google Scholar
Knauth, LP and Kennedy, MJ (2009) The late Precambrian greening of the earth. Nature 460, 728732.CrossRefGoogle Scholar
Knoll, AH, Bambach, RK, Canfield, DE and Grotzinger, JP (1996) Comparative earth history and late Permian mass extinction. Science 273, 452457.CrossRefGoogle ScholarPubMed
Knoll, AH and Carroll, SB (1999) Early animal evolution: Emerging views from comparative biology and geology. Science 284, 21292137.CrossRefGoogle ScholarPubMed
Knoll, AH and Walter, MR (1992) Latest Proterozoic stratigraphy and earth history. Nature 356, 673678.CrossRefGoogle ScholarPubMed
Kocsis, ÁT, Reddin, CJ, Alroy, J and Kiessling, W (2019) The R package divDyn for quantifying diversity dynamics using fossil sampling data. Methods in Ecology and Evolution 10, 735743.CrossRefGoogle Scholar
Kristensen, E, Penha-Lopes, G, Delefosse, M, Valdemarsen, T, Quintana, CO and Banta, GT (2012) What is bioturbation? The need for a precise definition for fauna in aquatic sciences. Marine Ecology Progress Series 446, 285302.CrossRefGoogle Scholar
Krug, AZ and Jablonski, D (2012) Long-term origination rates are reset only at mass extinctions. Geology 40, 731734.CrossRefGoogle Scholar
Ksepka, DT, Stidham, TA and Williamson, TE (2017) Early Paleocene landbird supports rapid phylogenetic and morphological diversification of crown birds after the K-Pg mass extinction. Proceedings of the National Academy of Sciences 114, 80478052.CrossRefGoogle ScholarPubMed
Laflamme, M, Darroch, SAF, Tweedt, SM, Peterson, KJ and Erwin, DH (2013) The end of the Ediacara biota: Extinction, biotic replacement, or Cheshire cat? Gondwana Research 23, 558573.CrossRefGoogle Scholar
Laflamme, M, Xiao, S and Kowalewski, M (2009) Osmotrophy in modular Ediacara organisms. Proceedings of the national Academy of Sciences 106, 1443814443.CrossRefGoogle ScholarPubMed
Leme, JM, Van Iten, H and Simoes, MG (2022) A new conulariid (Cnidaria, Scyphozoa) from the terminal Ediacaran of Brazil. Frontiers in Earth Science 10, 777746. https://doi.org/10.3389/feart.2022.777746.CrossRefGoogle Scholar
Li, Z, Cao, M, Lloyd, SJ, Algeo, TJ, Zhao, H, Wang, X, Zhao, L and Chen, Z-Q (2020) Transient and stepwise ocean oxygenation during the late Ediacaran Shuram excursion: Insights from carbonate δ238U of northwestern Mexico. Precambrian Research 344, 105741.CrossRefGoogle Scholar
Linnemann, U, Ovtcharova, M, Schaltegger, U, Gartner, A, Hautmann, M, Geyer, G, Vickers-Rich, P, Rich, T, Plessen, B, Hofmann, M, Zieger, J, Krause, R, Kriesfield, L and Smith, J (2019) New high-resolution age data from the Ediacaran–Cambrian boundary indicate rapid, ecologically driven onset of the Cambrian explosion. Terra Nova 31, 4958.CrossRefGoogle Scholar
Liu, AG, Kenchington, CG and Mitchell, EG (2015) Remarkable insights into the paleoecology of the Avalonian Ediacaran macrobiota. Gondwana Research 27, 13551380.CrossRefGoogle Scholar
Liu, AG, Matthews, JJ, Menon, LR, McIlroy, D and Brasier, MD (2014) Haootia quadriformis n. gen., n. sp., interpreted as a muscular cnidarian impression from the Late Ediacaran period (approx. 560 Ma). Proceedings of the Royal Society B 281, 20141202.CrossRefGoogle Scholar
Love, GD and Zumberge, JA (2021) Emerging patterns in Proterozoic lipid biomarkers: Implications for marine biospheric evolution and the ecological rise of eukaryotes. In Lyons, T, Turchyn, A and and Reinhard, C (eds.), Elements in Geochemical Tracers in Earth System Science. Cambridge: Cambridge University Press.Google Scholar
Lu, M, Lu, Y, Ikejiri, T, Sun, D, Carroll, R, Blair, EH, Algeo, TJ and Sun, Y (2021) Periodic oceanic euxinia and terrestrial fluxes linked to astronomical forcing during the late Devonian Frasnian–Famennian mass extinction. Earth and Planetary Science Letters 562, 116839.CrossRefGoogle Scholar
Lu, M, Zhu, M, Zhang, J, Shields-Zhou, G, Li, G, Zhao, F, Zhao, X and Zhao, M (2013) The DOUNCE event at the top of the Ediacaran Doushantuo Formation, South China: Broad stratigraphic occurrence and non-diagenetic origin. Precambrian Research 225, 86109.CrossRefGoogle Scholar
Macdonald, FA, Strauss, JV, Sperling, EA, Halverson, GP, Narbonne, GM, Johnston, DT, Kunzmann, M, Schrag, DP and Higgins, JA (2013) The stratigraphic relationship between the Shuram carbon isotope excursion, the oxygenation of Neoproterozoic oceans, and the first appearance of the Ediacara biota and bilaterian trace fossils in northwestern Canada. Chemical Geology 362, 250272.CrossRefGoogle Scholar
Maloof, AC, Porter, SM, Moore, JL, Dudas, FO, Bowring, SA, Higgins, JA, Fike, DA and Eddy, MP (2010) The earliest Cambrian record of animals and ocean geochemical change. GSA Bulletin 122, 17311774.CrossRefGoogle Scholar
Mángano, MG and Buatois, LA (2014) Decoupling of body-plan diversification and ecological structuring during the Ediacaran-Cambrian transition: Evolutionary and geobiological feedbacks. Proceedings of the Royal Society B 281, 20140038.CrossRefGoogle ScholarPubMed
Mángano, MG and Buatois, LA (2020) The rise and early evolution of animals: Where do we stand from a trace-fossil perspective? Interface Focus 10, 20190103.CrossRefGoogle ScholarPubMed
Marenco, KN and Bottjer, DJ (2007) Ecosystem engineering in the fossil record: Early examples from the Cambrian Period. In Cuddington, K, Byers, JE, Wilson, WG and Hastings, A (eds.), Ecosystem Engineers: Plants to Protists. Amsterdam: Academic Press, pp. 163183.CrossRefGoogle Scholar
Matthews, JJ, Liu, AG, Yang, C, McIlroy, D, Levell, B and Condon, DJ (2021) A Chronostratigraphic framework for the rise of the Ediacaran Macrobiota: new constraints from Mistaken Point Ecological Reserve, Newfoundland. Geological Society of America Bulletin 133, 612624.CrossRefGoogle Scholar
Meysman, FJR, Middelburg, JJ and Heip, CHR (2006) Bioturbation: A fresh look at Darwin’s last idea. Trends in Ecology and Evolution 21, 688695.CrossRefGoogle Scholar
Minter, NJ, Buatois, LA, Mangano, MG, Davies, NS, Gibling, MR, Macnaughton, RB and Labandeira, CC (2017) Early bursts of diversification defined the faunalcolonization of land. Nature Ecology and Evolution 1075.Google Scholar
Mitchell, EG, Evans, S, Chen, Z and Xiao, S (2022) A new approach for investigating spatial relationships of ichnofossils: A case study of Ediacaran–Cambrian animal traces. Paleobiology 48, 557575.CrossRefGoogle Scholar
Mitchell, EG, Harris, S, Kenchington, CG, Vixseboxse, P, Roberts, L, Clark, C, Dennis, A, Liu, AG and Wilby, PR (2019) The importance of neutral over niche processes in structuring Ediacaran early animal communities. Ecology Letters 22, 20282038.CrossRefGoogle ScholarPubMed
Muscente, AD, Bykova, N, Boag, TH, Buatois, LA, Mángano, GM, Eleish, A, Prabhu, A, Pan, F, Meyer, MB, Schiffbauer, JD, Fox, P, Hazen, RM and Knoll, AH (2019) Ediacaran biozones identified with network analysis provide evidence for pulsed extinctions of early complex life. Nature Communications 10, 911.CrossRefGoogle ScholarPubMed
Narbonne, GM (2005) The Ediacara biota: Neoproterozoic origin of animals and their ecosystems. Annual Reviews of Earth and Planetary Sciences 33, 421442.CrossRefGoogle Scholar
Narbonne, GM, Kaufman, AJ and Knoll, AH (1994) Integrated chemostratigraphy and biostratigraphy of the Windermere Supergroup, northwestern Canada: Implications for Neoproterozoic correlations and the early evolution of animals. Geological Society of America Bulletin 106, 12811292.2.3.CO;2>CrossRefGoogle Scholar
Nelson, LL, Ahm, A-SC, Macdonald, FA, Higgins, JA and Smith, EF (2021) Fingerprinting local controls on the Neoproterozoic carbon cycle with the isotopic record of Cryogenian carbonates in the Panamint Range, California. Earth and Planetary Science Letters 566, 116956.CrossRefGoogle Scholar
Nelson, LL, Ramezani, J, Almond, JE, Darroch, SAF, Taylor, WL, Brenner, DC, Furey, RP, Turner, M and Smith, EF (2022) Pushing the boundary: A calibrated Ediacaran–Cambrian stratigraphic record from the Nama group in northwestern Republic of South Africa. Earth and Planetary Science Letters 580, 117396.CrossRefGoogle Scholar
Newell, ND (1967) Revolutions in the history of life. Geological Society of America Special Publications 89, 6391.CrossRefGoogle Scholar
O’Leary, MA, Bloch, JI, Flynn, JJ, Gaudin, TJ, Giallombardo, A, Giannini, NP, Goldberg, SL, Kraatz, BP, Luo, Z-X, Meng, J, Ni, X, Novacek, MJ, Perini, FA, Randall, ZS, Rougier, GW, Sargis, EJ, Silcox, MT, Simmons, NB, Spaulding, M, Velazco, PM, Weksler, M, Wible, JR and Cirranello, AL (2013) The Placental Mammal Ancestor and the Post K-Pg Radiation of Placentals. Science 339, 662667.CrossRefGoogle ScholarPubMed
Oehlert, AM and Swart, PK (2014) Interpreting carbonate and organic carbon isotope covariance in the sedimentary record. Nature Communications 5, 4672.CrossRefGoogle ScholarPubMed
Ouyang, Q, Zhou, C, Xiao, S, Chen, Z and Shao, Y (2019) Acanthomorphic acritarchs from the Ediacaran Doushantuo formation at Zhangcunping in South China, with implications for the evolution of early Ediacaran eukaryotes. Precambrian Research 320, 171192.CrossRefGoogle Scholar
Park, T-YS, Jung, J, Lee, M, Lee, S, Zhen, YY, Hua, H, Warren, LV and Hughes, NC (2021) Enduring evolutionary embellishment of cloudinids in the Cambrian. Royal Society Open Science 8, 34909213.CrossRefGoogle ScholarPubMed
Parry, LA, Boggiani, PC, Condon, DJ, Garwood, RJ, Leme, JD-M, McIlroy, D, Brasier, MD, Trindade, R, Campanha, GAC, Pacheco, MLAF, Diniz, CQC and Liu, AG (2017) Ichnological evidence for meiofaunal bilaterians from the terminal Ediacaran and earliest Cambrian of Brazil. Nature Ecology and Evolution 10, 14551464. https://doi.org/10.1038/s41559-017-0301-9.CrossRefGoogle Scholar
Pineda-Munoz, S, Wang, Y, Lyons, SK, Toth, AB and McGuire, JL (2021) Mammal species occupy different climates following the expansion of human impacts. Proceedings of the National Academy of Sciences 118, e1922859118.CrossRefGoogle ScholarPubMed
Pu, JP, Bowring, SA, Ramezani, J, Myrow, P, Raub, TD, Landing, E, Mills, A, Hodgin, E and Macdonald, FA (2016) Dodging snowballs: Geochronology of the Gaskiers glaciation and the first appearance of the Ediacaran biota. Geology 44, 955958.CrossRefGoogle Scholar
Rahman, IA, Darroch, SAF, Racicot, RA and Laflamme, M (2015) Suspension feeding in the enigmatic Ediacaran organism Tribrachidium demonstrates complexity of Neoproterozoic ecosystems. Science Advances 1, e1500800.CrossRefGoogle ScholarPubMed
Raup, DM and Sepkoski, JJ (1982) Mass extinctions in the marine fossil record. Science 215, 15011503.CrossRefGoogle Scholar
Rhoads, DC, Yingst, JY and Ullman, WJ (1978) Seafloor stability in Central Long Island sound: Part I. temporal changes in erodibility of fine-grained sediment. In Wiley, ML (ed.), Estuarine Interactions. New York: Academic Press, pp. 221244.CrossRefGoogle Scholar
Rooney, AD, Cantine, MD, Bergmann, KD, Gomez-Perez, I, Al Baloushi, B, Boag, TH, Busch, JF, Sperling, EA and Strauss, JV (2020) Calibrating the coevolution of Ediacaran life and environment. Proceedings of the National Academy of Sciences USA 117, 1682416830.CrossRefGoogle ScholarPubMed
Rosenberg, R, Nilsson, HC and Diaz, RJ (2001) Response of benthic fauna and changing sediment redox profiles over a hypoxic gradient. Estuarine and Coastal Shelf Science 53, 343350.CrossRefGoogle Scholar
Ruhl, M, Bonis, NR, Reichert, G-J, Damste, JSS and Kurschner, WM (2011) Atmospheric carbon injection linked to end-Triassic mass extinction. Science 333, 430434.CrossRefGoogle ScholarPubMed
Schiffbauer, JD, Huntley, JW, O’Neil, GR, Darroch, SAF, Laflamme, M and Cai, Y (2016) The latest Ediacaran wormworld fauna: Setting the ecological stage for the Cambrian explosion. GSA Today 26, 411.CrossRefGoogle Scholar
Schiffbauer, JD, Rosbach, S, Pulsipher, MA, Leibach, W, Nolan, M, Tang, Q, Lindsay-Kauffman, A and Selly, T (2022). Diversity in the age of tubes: A morphometric approach to understanding the breadth of Ediacaran tube-dwelling taxa. Geological Society of America Abstracts with Programs 54. doi: 10.1130/abs/2022AM-381858CrossRefGoogle Scholar
Schiffbauer, JD, Selly, T, Jacquet, SM, Merz, RA, Nelson, LL, Strange, MA, Cai, Y and Smith, EF (2020) Discovery of bilaterian-type through-guts in cloudinomorphs from the terminal Ediacaran Period: Nature Communications 11, 112.Google Scholar
Seilacher, A (1984) Late Precambrian and early Cambrian Metazoa: Preservational or real extinctions? In Holland, HD and Trendall, AF (eds.), Patterns of Change in Earth Evolution. Heidelberg, Germany: Springer-Verlag, pp. 159168.CrossRefGoogle Scholar
Seilacher, A (1985) Discussion of Precambrian metazoans. Philosophical Transactions of the Royal Society of London Series B 311, 4748.Google Scholar
Seilacher, A (1989) Vendozoa: Organismic construction in the Proterozoic biosphere. Lethaia 22, 229239.CrossRefGoogle Scholar
Seilacher, A (1992) Vendobionta and Psammocorallia: Lost constructions of Precambrian evolution. Journal of the Geological Society of London 149, 607613.CrossRefGoogle Scholar
Shen, SZ, Crowley, JL, Wang, Y, Bowring, SA, Erwin, DH, Sadler, PM, Cao, C-Q, Rothman, DH, Henderson, CM, Ramezani, J, Zhang, H, Shen, Y, Wang, X-D, Wang, W, Mu, L, Li, Y-G, Liu, Z-L, Liu, L-J, Zheng, Y, Y-F, J and Jin, Y-G (2011) Calibrating the end-Permian mass extinction. Science 334, 13671372.CrossRefGoogle ScholarPubMed
Shore, AJ, Wood, RA, Butler, IB, Zhuravelev, AY, McMahon, S, Curtis, A and Bowyer, FT (2021) Ediacaran metazoan reveals lophotrochozoan affinity and deepens root of Cambrian explosion. Science Advances 7, eabf2933. https://doi.org/10.1126/sciadv.abf293.CrossRefGoogle ScholarPubMed
Shore, AJ, Wood, RA, Curtis, A and Bowyer, FT (2020) Multiple branching and attachment structures in cloudinomorphs, Nama Group, Namibia. Geology 48, 877881.CrossRefGoogle Scholar
Slagter, S, Hao, W, Planavsky, NJ, Konhauser, KO and Tarhan, LG (2022) Biofilms and agents of Ediacara-style fossilization. Scientific Reports 12, 8631.CrossRefGoogle ScholarPubMed
Smith, EF, Nelson, LL, O’Connell, N, Eyster, AE and Lonsdale, MC (2023) The Ediacaran− Cambrian transition in the southern Great Basin, United States. GSA Bulletin 135, 13931414.Google Scholar
Smith, EF, Nelson, LL, Strange, MA, Eyster, AE, Rowland, SM, Schrag, DP and Macdonald, FA (2016) The end of the Ediacaran: Two exceptionally preserved body fossil assemblages from Mount Dunfee, Nevada, USA. Geology 44, 911914.CrossRefGoogle Scholar
Swart, P (2008) Global synchronous changes in the carbon isotopic composition of carbonate sediments unrelated to changes in the global carbon cycle. Proceedings of the National Academy of Sciences USA 105, 1374113745.CrossRefGoogle ScholarPubMed
Tarhan, LG, Droser, ML, Cole, DB and Gehling, JG (2018) Ecological expansion and extinction in the late Ediacaran: Weighing the evidence for environmental and biotic drivers. Integrative and Comparative Biology 58, 688702.CrossRefGoogle ScholarPubMed
Tarhan, LG, Droser, ML, Planavsky, NJ and Johnston, DT (2015) Protracted development of bioturbation through the early Palaeozoic era. Nature Geoscience 8, 865869.CrossRefGoogle Scholar
Tarhan, LG, Zhao, M and Planavsky, NJ (2021) Bioturbation feedbacks on the phosphorous cycle. Earth and Planetary Science Letters 556, 116961.CrossRefGoogle Scholar
Topper, T, Betts, MJ, Dorjnamjaa, D, Li, G, Li, L, Altanshagai, G, Enkhbaatar, B and Skovsted, CB (2022) Locating the BACE of the Cambrian: Bayan Gol in southwestern Mongolia and global correlation of the Ediacaran–Cambrian boundary. Earth Science Reviews 226, 104017.CrossRefGoogle Scholar
Turk, KA, Maloney, KM, Laflamme, M and Darroch, SAF (2022) Paleontology and ichnology of the late Ediacaran Nasep-Huns transition (Nama group, southern Namibia). Journal of Paleontology 96, 753769.CrossRefGoogle Scholar
Verdel, C, Wernicke, BP and Bowring, SA (2011) The Shuram and subsequent Ediacaran carbon isotope excursions from southwest Laurentia, and implications for environmental stability during the metazoan radiation. Geological Society of America Bulletin 123, 15391559.CrossRefGoogle Scholar
Waggoner, B (1999) Biogeographic analyses of the Ediacara biota: A conflict with paleotectonic reconstructions. Paleobiology 25, 440458.CrossRefGoogle Scholar
Waggoner, B (2003) The Ediacaran biotas in space and time. Integrative and Comparative Biology 43, 104113.CrossRefGoogle ScholarPubMed
Wang, X, Chen, Z, Pang, K, Chuanming, Z, Xiao, S, Wan, B and Yuan, X (2021) Dickinsonia from the Ediacaran Dengying formation in the Yangtze gorges area, South China. Palaeoworld 30, 602609.CrossRefGoogle Scholar
Wood, RA and Curtis, A (2014) Extensive metazoan reefs from the Ediacaran Nama Group, Namibia: the rise of benthic suspension feeding. Geobiology 13, 112122.CrossRefGoogle ScholarPubMed
Wood, RA, Liu, AG, Bowyer, F, Wilby, PR, Dunn, FS, Kenchington, CG, Hoyal-Cuthill, JF, Mitchell, EG and Penny, A (2019) Integrated records of environmental change and evolution challenge the Cambrian explosion. Nature Ecology and Evolution 3, 528538.CrossRefGoogle ScholarPubMed
Wood, RA, Poulton, SW, Prave, AR, Hoffmann, K-H, Clarkson, MO, Guilbaud, R, Lyne, JW, Tostevin, R, Bowyer, F, Penny, AM, Curtis, A and Kasemann, SA (2015) Dynamic redox conditions control late Ediacaran metazoan ecosystems in the Nama group, Namibia. Precambrian Research 261, 252271.CrossRefGoogle Scholar
Xiao, S, Narbonne, GM, Zhou, C, Laflamme, M, Grazhdankin, D, Moczydlowska-Vidal, M and Cui, H (2016) Towards an Ediacaran time scale: problems, protocols, and prospects. Episodes 36, 540555.CrossRefGoogle Scholar
Xiao, S, Chen, Z, Pang, K, Zhou, C and Yuan, X (2020) The Shibantan Lagerstätte: Insights into the Proterozoic–Phanerozoic transition. Journal of the Geological Society 178, jgs2020–135.Google Scholar
Yang, B, Steiner, M, Schiffbauer, JD, Selly, T, Wu, X, Zhang, C and Liu, P (2020) Ultrastructure of Ediacaran cloudinids suggests diverse taphonomic histories and affinities with non-biomineralized annelids. Scientific Reports 10, 535.CrossRefGoogle ScholarPubMed
Yang, B, Steiner, M, Zhu, M, Guoxiang, L, Liu, J and Liu, P (2016) Transitional Ediacaran–Cambrian small skeletal fossil assemblages from South China and Kazakhstan: implications for chronostratigraphy and metazoan evolution. Precambrian Research 285, 202215.CrossRefGoogle Scholar
Yang, C, Rooney, AD, Condon, DJ, Li, X-H, Grazhdankin, DV, Bowyer, FT, Hu, C, Macdonald, FA and Zhu, M (2021) The tempo of Ediacaran evolution. Science Advances 7, eabi9643.CrossRefGoogle ScholarPubMed
Yeakel, JD, Pires, MM, de Aguiar, MAM, O’Donnell, JL, Guimaraes, PR, Gravel, D and and Gross, T (2020) Diverse interactions and ecosystem engineering can stabilize community assembly. Nature Communications 11, 3307.CrossRefGoogle ScholarPubMed
Youbi, N, Ernst, RE, Söderlund, U, Boumehdi, MA, Lahna, A, Tassinari, CCG, El Moume, W and Bensalah, MK (2020) The Central Iapetus magmatic province (CIMP): An updated review and link with the c. 580 ma Gaskiers glaciation. In Adatte, T., Bond, D.P.G., and Keller, G., eds. New Developments on Volcanism, Impacts and Mass Extinctions. Geological Society of America Special Paper 544.Google Scholar
Zhang, F, Xiao, S, Romaniello, SJ, Hardisty, D, Li, C, Melezhik, V, Pokrovsky, B, Cheng, M, Shi, W, Lenton, TM and Anbar, AD (2019) Global marine redox changes drove the rise and fall of the Ediacara biota. Geobiology 17, 594610.CrossRefGoogle ScholarPubMed
Zhu, M, Zhuravlev, AY, Wood, RA, Zhao, F and Sukhov, SS (2017) A deep root for the Cambrian explosion: implications of a new bio- and chemostratigraphy from the Siberian Platform. Geology 45, 459462.CrossRefGoogle Scholar
Figure 0

Figure 1. Updated summary figure illustrating the stratigraphic distribution and diversity among groups of Ediacara biota, as well as metazoans, bilaterian ichnogenera, and a δ13C curve (compiled from Yang et al. (2021), Bowyer et al. (2022), and references therein). The stratigraphic ranges of the Pentaradialomorpha and Tetraradialomorpha are currently uncertain, but currently constrained by a detrital zircon age of 556 ± 24 Ma obtained from the Bonney Sandstone in South Australia (Ireland et al., 1998). Solid colors represent minimum age estimates (where available), while shaded regions represent uncertain range estimates where taxa are found beneath (or between) dated horizons. Extinction intensities – as percentage of genera lost – are given for the two putative extinction pulses at the White Sea-Nama and the E–C boundaries; intensities were calculated by simply measuring the proportion of surviving genera over total genera in the preceding assemblage zone (although see discussion in the text surrounding problems with calculating these transition).

Figure 1

Figure 2. Putative late Ediacaran ecosystem engineers, including bilaterian tracemaking behaviors that involve sediment ‘bulldozing’ and biomixing (A - Parapsammichnites), bioirrigation (B-C - large treptichnids), and suspension feeders such as Paleophragmodictya (D-E; sp. nos. P32338 and P32332-P32352 respectively, South Australia Museum), biomineralizing Cloudina (F), and other unidentified tubefauna (G). Lastly, many Ediacara biota may have also had important ecosystem engineering impacts; the enigmatic taxa Ernietta (J), Arkarua (H; sp. no. P26768, South Australia Museum), Tribrachidium (I; sp. no. N3993/5056, Palaeontological Institute, Moscow) and Pteridinium (K) are all also thought to have functioned as suspension feeders, and thus played a crucial role in forging energetic links between the pelagic and benthic realms (Cracknell et al., 2021; Darroch et al., 2022). Specimens shown in A–C, F, G, J and K from the Nama Group of southern Namibia (all Urusis Fm., with the exception of Ernietta shown in J from the Dabis Fm.), and photographed in the field. Filled scale bars = 1 cm, open scale bars = 5 mm.

Figure 2

Figure 3. Ediacaran–Cambrian rift volcanism in Laurentia and timing of interpreted rifts leading to passive margin development around Laurentia after the breakup of Rodinia. Rifting along the southwestern margin has recently been suggested to coincide with the E–C boundary, potentially resulting in the BACE negative carbon isotope excursion and a second pulse of Ediacaran extinction (see Hodgin et al., 2021; Smith et al., 2022). CAMP, Central Atlantic magmatic province; CIMP, Central Iapetus magmatic province; Panth, Panthalassa Ocean.

Figure 3

Figure 4. A hypothetical model for drivers of the E–C transition, modeled after Knoll and Carroll (1999); their figure 5). ‘AV’ = Avalon; ‘WS’ = White Sea; ‘NAM’ = Nama. This model interprets the majority of ‘Ediacara biota’ as stem-group metazoans, with a first pulse of extinction at the White Sea-Nama transition driven by the diversification – with associated downstream geobiological impacts – of crown-group metazoan clades (i.e., ‘biotic replacement’). A second pulse of extinction follows at the E–C boundary, driven by widespread environmental perturbation (i.e., ‘catastrophe’) following extensive rift volcanism. We note that this figure is strictly hypothetical, and the stem – and crown-groups depicted here do not intentionally correspond to specific biological groups shown in Figure 1.

Supplementary material: File

Darroch et al. supplementary material

Darroch et al. supplementary material

Download Darroch et al. supplementary material(File)
File 18.8 KB

Author comment: Causes and consequences of end-Ediacaran extinction: An update — R0/PR1

Comments

Dear Editors,

On behalf of myself and co-authors, I am pleased to submit our manuscript, “Causes and consequences of end-Ediacaran extinction – an update” to be considered for publication in Cambridge Prisms: Extinction. This is an invited paper originally solicited by Dr. Wolfgang Kiessling (FAU Erlangen).

The Ediacaran–Cambrian (‘E-C’) transition marks one of the most important geobiological revolutions in Earth History, including multiple waves of evolutionary radiation, and – potentially – successive episodes of extinction. Discussion surrounding these events has intensified in the last decade, in concert with efforts to understand drivers of global change over the E-C boundary and the appearance of the more modern-looking Phanerozoic biosphere. Five years ago, several of these same authors summarized the evidence for one or more pulses of extinction in the late Ediacaran, and presented a series of key questions that would be crucial for driving knowledge forward in this field. Since then many of these questions have been explored, with new paleontological, geochemical, and geochronological datasets providing the scaffolding required for building our understanding of this interval. In this review we summarize recent work surrounding the end-Ediacaran extinction events, examine the extent to which the questions we posed in 2018 have been answered, and propose new questions and research avenues that will continue to illuminate the profound changes that occurred over this interval. Lastly, we propose a revised model for drivers of late Ediacaran extinction pulses that builds off recent data and growing consensus within the field. This model is speculative, but does frame testable hypotheses that can be targeted in the next decade of work.

This material has not been submitted for publication elsewhere, and will not be submitted while it is in review.

Yours sincerely,

Simon Darroch

Vanderbilt University

Department of Earth and Environmental Sciences

Review: Causes and consequences of end-Ediacaran extinction: An update — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

Unsurprisingly, this manuscript is about extinction, which is the ultimate fate of all taxa, just as death comes eventually to all individuals. So the question is, were the extinctions of Ediacaran taxa synchronous enough to move a normal process into the abnormal range of mass extinction and thus warrant this protracted discussion?

Here, the reliability of the observed fossil record becomes paramount. Are the Avalon, White Sea and Nama assemblages really separated by mass extinctions or are they windows into an evolving biota where change seems abrupt because of intermittent viewing? It is circular reasoning to think that the White Sea and undated South Australian assemblages are the same ago because they so similar, and then to use the fact that they are so similar to argue that the known duration of the White Sea assemblages, based on U-Pb ages, implies that the South Australian examples went extinct at the same time.

Furthermore, the Nama assemblage consists of a few distinctive taxa (Pteridinium, Rangea, Archaeichnium, Ernietta, Swartpuntia) that—at best—are known from rare specimens elsewhere. In South Australia, where the succession is clear and the sampling outstanding, the Nama taxa Pteridinium and Rangea are only found beneath the beds that contain the ‘White Sea’ assemblage despite almost a century of targeted collecting.

Line 49: 1980s not 1980’s.

121-123: Whether the Ediacara biota were metazoans or not has no bearing on their propensity for extinction. In any case, the Ediacara biota has already been segregated from the Metazoa (102) because they lack the body plans of the animal phyla.

133-136: There has been a lot of water under the bridge since Amthor et al (2003) and the results they presented are probably incorrect about the age of BACE and the extinction of cloudinids. In any case, losing a couple of calcareous taxa (Cloudina and Namacalathus) is hardly grounds for a mass extinction. The authors seem to confuse fossil abundance with taxonomic diversity.

137-153: Everyone agrees that there are more Ediacara biota taxa in the White Sea assemblages than in the Nama assemblage, but is this because of extinction or loss of information? Apart from the famous Pteridinium locality at Aar farm, Ediacaran fossils are rare in Namibia and only one or two modes of preservation are found. Most importantly, there are very few of the fine bed surfaces that are so common at numerous White Sea and South Australian localities. With this background, it is not helpful to use the ‘Kotlinian crisis’, which - as the authors acknowledge - was introduced for a fundamentally different reason to support their purported White Sea-Nama mass extinction event.

174-176: ‘genus richness among the erniettomorphs etc.’ This doubtful argument is based on a small number of rare fossils (both taxa and specimens). Ernietta is known from only two narrow stratigraphic intervals, one in Namibia and one in Nevada. Swartpuntia, another genus that increases the biodiversity of the Nama assemblage, perhaps enough to tip the balance, is known from only one place on the planet if doubtful identifications in Nevada, Australia and North Carolina are disregarded, as they should be for discussions of this nature.

190-198: The evidence presented for extinction at the Ediacaran-Cambrian boundary suffers from another difficulty: taxonomic splitting. It is indisputable that ‘worm tubes’ and cloudinids are common fossils in the latest Ediacran, but anyone who has worked on small shelly fossils from the early Cambrian would know that tube fossils did not go extinct. They just became less important by dilution with other more interesting, shelly fossils. Because tubular fossils are so rare in the Proterozoic they have been studied and named far more thoroughly, so counting the loss of generic names across the boundary is not necessarily good evidence for mass extinction.

266: ‘followed by the origin of new clades’ is not correct. Most post-extinction expansive clades (angiosperms, birds, mammals, ammonites) originated prior to the extinction event. They may have radiated spectacularly afterwards but they did not originate after the event.

313-314: Add a reference to Topper et al. (2022) when discussing the timing of the BACE.

Fig. 1. This figure is misleading in that it implies that the sudden originations and expansions at the beginnings of the periods represented by Avalon, White Sea and Nama are based on evidence from the fossil record. There is no evidence for synchronous origins (or extinctions) of these putative clades. Their stratigraphic ranges should be shown as bars with fuzzy beginnings and endings. Leaving out the horizontal lines, which imply a precision that does not exist, might help. For example, even if Kimberella is a bilaterian - which many dispute - there is no evidence that its origin coincided with the first appearance of other members of the White Sea assemblage. Trace fossils may be a more useful guide to the time of appearance of bilaterians. Similarly, the presence of Porifera and Cnidaria in the Avalon assemblage is by no means certain, so it is even more problematical to postulate that both phyla appeared synchronously at the beginning of Avalon time. If these uncertainties are properly displayed, the case for mass extinction can more easily be evaluated by the readers. Also, Porifera (spicular sponges and archaeocyaths) underwent a spectacular radiation during the early Cambrian, not shown on this figure.

Fig. 4. Avoid following Knoll and Carroll (1999) for the model shown in Figure 4. There has been a huge increase in knowledge from genomics since then and most biologists would now place sponges and their grade of organization at the base of the metazoan tree. If so, almost all members of the Ediacara biota become crown not stem metazoans, as discussed earlier in the manuscript. This problem could be finessed in Fig. 4 by speaking about stem and crown eumetazoans, but the branch order in the stem group then needs to be addressed and made congruent with the clades shown in Fig. 1. At present, these two figures present very different views of the Ediacaran history of the Metazoa. In Fig. 1, the metazoan crown group originates prior to Avalon time, the Bilateria and presumably the Lophotrochozoa prior to White Sea time, and six Ediacara Biota/stem Eumetazoa groups originate synchronously at the Avalon-White Sea boundary, as shown by their point sources.

Also, where are the data that would support a bottleneck for the crown taxa at the E-C boundary? The worm tubes? Cloudina (arguably a metazoan) and Namacalathus (probably not)? According to Fig. 1 all metazoan groups except worm tubes, which in any case are only form taxa, were diversifying at this time.

In summary, the case for mass extinctions at the end of White Sea time and at the end of the Ediacaran is not strong. Although widespread negative carbon isotope excursions (BANE and BACE) may have occurred at the same times, there is no evidence apart from temporal coincidence, that these excursions were involved in the proposed extinction events. There is a need to address the discrepancies between Figs. 1 and 4. Fig. 3 could be omitted as it is barely discussed and only applies to Laurentia.

Review: Causes and consequences of end-Ediacaran extinction: An update — R0/PR3

Conflict of interest statement

Reviewer declares none.

Comments

This manuscript reviews current thinking and recent research regarding proposed mass extinction events in the latest Ediacaran Period, before posing a series of new questions that could provide a framework for future research in this area. It is a well written and coherent summary of this topic, and addresses several pertinent issues. There are undoubtedly other questions that could be considered, but the article works as a succinct and approachable overview of current thinking. I have suggested a few places where additional references could be considered, and one potential edit to Figure 4, but I have no major concerns, and I congratulate the authors on producing an enjoyable and informative paper.

The authors may want to consider discussion/inclusion of the recent Evans et al (2022) paper in PNAS, published after submission of this manuscript:

Evans, Scott D., et al. “Environmental drivers of the first major animal extinction across the Ediacaran White Sea-Nama transition.” Proceedings of the National Academy of Sciences 119 (2022): e2207475119.

Minor comments:

l. 96: There is scope to refer to biomarker evidence for metazoans here too (Bobrovskiy et al 2018 Science paper on Dickinsonia, or their very recent 2022 Current Biology paper on Kimberella and Calyptrina diet). It might also be relevant to mention the trace fossil record early here, to demonstrate the co-existence of latest Ediacaran biotas with bilaterian trace makers.

l. 153: The “Kotlinian” is time-equivalent to this crisis, as discussed in Grazhdankin (2014) (Patterns of evolution of the Ediacaran soft-bodied biota. Journal of Paleontology, 88(2), 269-283).

l. 166-169: This statement seems quite ‘certain’, given the uncertainty surrounding timings and ages of many global sites (e.g. South Australia), and recent discoveries (such as dickinsoniid specimens in the ?Nama-aged Shibantan Mbr: Wang et al., (2021), as you mention later).

l. 198: If mentioning the possibility of survivorship of Ediacaran taxa across the boundary, it is probably worth noting the existence of candidate soft-bodied ‘survivors’ too for completeness (recently summarised in Hoyal Cuthill, 2022: Ediacaran survivors in the Cambrian: suspicions, denials and a smoking gun. Geological Magazine, 1-10; or this very recent paper: Hu et al (2022). A new Cambrian frondose organism:“ Ediacaran survivor” or convergent evolution?. Journal of the Geological Society, jgs2022-088).

l. 248: Perhaps cite some additional examples for these points (as you have done for the Shuram above).

l. 261-263; Arguably there was/is minimal independent geological evidence for a catastrophe either (in the form of LIPs, impacts, OAEs or any of the other usual suspects correlative particularly with the older event).

l. 288: Matthews et al 2021, not 2020

l. 366: Mangano and Buatois 2020 also seems relevant as a review of this extended trace fossil record here: Mángano, M. G., & Buatois, L. A. (2020). The rise and early evolution of animals: where do we stand from a trace-fossil perspective?. Interface Focus, 10(4), 20190103.

l. 407: This is a valuable point, but it would be more informative to a reader if the rationale provided for those counter-claims by Wood et al (2019) were summarised here, and then elaborated on to present how, if at all, those alternative interpretations feed into the question of whether the extinction pulses were rapid or slow. Unfortunately I fear that even if the Shibantan presents an opportunity to track evolution across this interval, determining whether the transition is fast or slow will require high-precision dating within that succession, which may be difficult/impossible to achieve without methodological advances.

l. 435: On this point, better constraint of palaeolatitudinal characteristics would also be welcome, and has recently been attempted by both Evans et al 2022 (above), and Boddy et al 2022: Palaeolatitudinal distribution of the Ediacaran macrobiota. Journal of the Geological Society, 179(1).)

l. 445: Jensen et al 1998 discuss their specimens as ‘Swartpuntia-like’ rather than arboreomorphs. I’m not aware of any reassessment of the specimens since? If I’m correct here, Figure 1 may require editing.

l. 479: Studies such as Lu et al (2013; The DOUNCE event at the top of the Ediacaran Doushantuo Formation, South China: Broad stratigraphic occurrence and non-diagenetic origin. Precambrian Research, 225, 86-109.) demonstrate a facies dependence for the nature of CIEs in the Neoproterozoic too, which is of relevance here.

l. 514: ‘treptichnids record…’? Rather than treptichnids are.

l. 535 (see also l. 548): Discussion of some of the work Emily Mitchell is doing on palaeoecology might be relevant here (e.g. Mitchell et al 2019, The importance of neutral over niche processes in structuring Ediacaran early animal communities. Ecology letters, 22(12), 2028-2038, or her recent trace fossil paper Mitchell, E. G., Evans, S. D., Chen, Z., & Xiao, S. (2022). A new approach for investigating spatial relationships of ichnofossils: a case study of Ediacaran–Cambrian animal traces. Paleobiology, 1-19.)

l. 966: How were these extinction intensities calculated?

Figure 1: I’m interested to know more about the basis of the data behind the shapes of the Cnidaria and Porifera diversity curves here (what exactly are the sponges in the White Sea assemblage, and do these plots account for basal/early Cambrian sponge taxa discovered from the Yanjiahe Fm, Soltanieh Fm, or the Hetang biota?) Which taxa are being interpreted as cnidarians in the Nama and earliest Cambrian settings? (and if you’re referring to tubular taxa as candidate cnidarians, are they being counted twice here?)

Figure 2 caption: please include information on where these specimens are from (locality and stratigraphy), along with any associated museum accession numbers (the Australian specimens in particular should include any SAM P numbers).

Figure 4: This figure works as a simple conceptual model, but an alternative version could envisage the Ediacaran biota including stem group members of multiple extant metazoan (and algal/protistan) clades that had already diverged prior to either extinction pulse. So, rather than one lineage crossing the boundary, I would envisage multiple lineages crossing the boundary, each with their own stem groups (e.g. what you’ve drawn being repeated for cnidarians, sponges, and bilaterians at least as individual clades within the Metazoa, in addition to algae and others. Potential stem group metazoans, stem group eumetazoans, and stem groups to individual metazoan phyla could also be truncated by the extinction events, or cross them). This adds complexity, but I think is more phylogenetically correct given current understanding.

Review: Causes and consequences of end-Ediacaran extinction: An update — R0/PR4

Conflict of interest statement

I know some of the co-authors, but this did not interfere with my objective review of the manuscript.

Comments

I enjoyed reading the manuscript ‘Causes and consequences of end-Ediacaran extinction – an update’, submitted to Cambridge Prisms: Extinction by Darroch et al. I think that the paper will be a valuable resource and framework for future work on the Ediacaran–Cambrian transition. Please find my review below:

This review-style paper of the Ediacaran-Cambrian transition serves as a 5-year update of Darroch et al. (2015). There are many new and interesting discussion points, in particular about patterns of extinction selectivity. The synthesis is valuable in re-evaluating the possibility of extinction occurring in the late Ediacaran and at the Ediacaran-Cambrian boundary. The review is prescient given that it has recently been shown that there is considerable uncertainty in the timing of the Ediacaran-Cambrian transition. The paradigm of two extinctions with the first driven by biotic replacement and the latter driven by environmental perturbation is a useful framework. The review paper serves as a fresh forward-looking synthesis that can help enliven debate in upcoming years, and serve as a reference point for considering the role of extinction events leading up to the ‘Cambrian Explosion’. Minor revisions are suggested for the text and figures. The minor revision suggestions can be found below amongst the line edits:

Line 55 — …several of the authors listed…

Lines 66-77 — consider adding references when introducing terminology such as “Rise of animals” or the “Big 5”

Line 157 — add a period

Line 189 — Following this summary, perhaps make a summary statement about the likelihood of late Ediacaran extinctions events.

Lines 231-41 — This may be one of the key questions posed in the 2018 paper, yet it is not clearly motivated here. It is not immediately obvious why the Shuram excursion is relevant to End-Ediacaran extinction? Perhaps an additional sentence of introduction/motivation would be helpful to map out the linkage (e.g., explicitly discussing its potential relationship to an origin of Ediacara biota or to an Ediacaran extinction event or that it could be an analogue to the BACE due to the magnitude of the excursion?).

Lines 284-88 — There are other excursions between the Shuram and the BACE… okay, I see this is mentioned in the next paragraph.

Lines 303-04 — It may be worth mentioning Verdel et al (2011) and the excursion in the Stirling Quartzite.

Lines 340 — Given the age constraints from Nelson et al (2022) and the suggestion that the BACE and the basal Cambrian may <538 Ma, it could be worth adding a sentence about the younger ages of the more voluminous portions of the Wichita Large Igneous Province.

Line 469 — Change ‘or’ to ‘of’

Line 485 — negative carbon isotope excursions

Line 498 — across the E–C transition.

Line 519 — no hyphen needed, in vivo

Line 576 — Earth history (no capitalization necessary of history)

Lines 585-94 — Consider making this one sentence using semi-colons (…specifically: (1) ... ; (2) … ; (3).)

Line 589 — Can the White Sea-Nama transition be considered a boundary?

Line 589 — ‘pulse of extinction is recognized at the White Sea-Nama boundary

Line 594 — There is a lot of information within the three-point summary. Prior to the final sentence inserted on line 594, it might be good to add a more condensed summary in which you refer to Fig. 4. Maybe something like: “To summarize, there are two apparent stages of extinction with the first driven by biotic replacement and the latter driven by environmental perturbation (Fig. 4).”

It might be worth reinforcing key concepts brought up in the introduction, such as extinction leading up to and at the E-C boundary being potentially classified along with the ‘Big 5’.

Line 966 — % age, why not write out percentage?

Line 971 — bioirrigation

Line 983 — Why is the Columbia River Flood Basalt included in the figure? It isn’t obvious that it’s relevant as point of comparison, being unrelated to rifting or extinction and being temporally far-removed from the late Ediacaran to E-C boundary.

Figures — Only figures 1 and 3 are referred to in the main text.

Figure 3 — i) On the NW margin of Laurentia, what is the timing of rifting referred to as “Ediacaran-Cambrian rifting” when numeric age ranges are given for the other margins?

ii) There is some agreement in color between panel a and panel b, such as purple = CIMP, green = Wichita. Should there be some sort of legend or explanation for the colors on the rift-related colors on the western and northwestern margins?

iii) Why is the CRFB shown? It does not appear on the map. It doesn’t appear to serve a purpose and it is not referred to in the main text.

Figure 4 — Explain what AV, WS, and NAM stand for.

Recommendation: Causes and consequences of end-Ediacaran extinction: An update — R0/PR5

Comments

Dear Authors,

Thank you for submitting your manuscript to our journal. Based on the evaluations provided by three expert reviewers and my own reading of the manuscript, your submission requires substantial revisions before it can be accepted for publication. The most critical review (’Major Revisions‘) raised multiple important points that need to be addressed. In addition, the other two reviewers -- despite recommending ’Minor Revisions‘ and ’Accept', respectively -- provided numerous useful suggestions that should be addressed carefully as well.

When returning the revised manuscript, please make sure to provide a detailed rebuttal letter explaining all your responses to the reviewers' comments.

Thank you for submitting your work to our journal,

Michal Kowalewski

Decision: Causes and consequences of end-Ediacaran extinction: An update — R0/PR6

Comments

No accompanying comment.

Author comment: Causes and consequences of end-Ediacaran extinction: An update — R1/PR7

Comments

To the editorial board at Cambridge Prisms: Extinction,

On behalf of myself and co-authors, we are pleased to re-submit our revised manuscript, “Causes and consequences of end-Ediacaran extinction – an update” to be considered for publication in Cambridge Prisms. This is an invited paper originally solicited by Dr. Wolfgang Kiessling (FAU Erlangen).

We would like to extend heartfelt thanks to our reviewers for their detailed and constructive comments – we were excited to read their responses to our work, and we have made the vast majority of suggested revisions. Specifically, we have substantially revised our Figures 1 and 4 to better match the fossil data and current phylogenetic inference (something requested by both Reviewers 2 and 3). We have also added a number of new recent references (Reviewer #2), and expanded the text in the Summary section to provide a more comprehensive – but still succinct – account of important take-home messages (Reviewer #3). We found comments provided by Reviewer #1 to be harder to manage; although we value their critical approach, this reviewer expressed several confusing and demonstrably incorrect views on the Ediacaran fossil record that suggests that they have not been keeping abreast with recent work on the interval, or otherwise have biased and/or fringe opinions of the evidence presented. We point in particular to statements like ‘Nama-aged fossils are rare and poorly-preserved’ (comments #1, 5 and 6 in the attached response-to-reviews document) – something that is easily falsified by a cursory reading of the last 10 years of research from multiple independent research groups working in Namibia, China, Iran, western USA, and South America. Another such comment is: ‘many people doubt Kimberella is a bilaterian’ (comment #10); to the best of our knowledge there have been no credible published papers opposing a bilaterian affinity since the late 1990s. Despite these complications, we have taken this review seriously and have added a new section to the manuscript titled: “Do these intervals of biotic turnover represent extinctions at all?” that tackles the central thesis put forward by Reviewer #1 – that the extinction events are an artifact of stratigraphy and sampling. This section unfortunately re-treads ground covered in many other papers written on the Ediacaran-Cambrian transition in the last 10 years, but may be justified in order to make this review more accessible to people less familiar with more recent work. Adding this section has made the manuscript ~1000 words longer, but we have done our best to consolidate text in other places to keep the piece succinct and readable.

A comprehensive list of our responses to reviewers’ comments is provided in our attached response document, along with a detailed list of changes made. For both clarity and ease of reference, we have highlighted reviewer comments in blue text. Our responses and details of amended analyses and text are given in black. Where appropriate, I have copied the amended text below our responses in italics, and with line references.

Should you have any additional questions, please do not hesitate to contact me.

Yours sincerely,

Simon Darroch (Vanderbilt University)

Review: Causes and consequences of end-Ediacaran extinction: An update — R1/PR8

Conflict of interest statement

Reviewer declares none.

Comments

Reviewer 1 is still not convinced, even with an adequate knowledge of the recent literature. However, as this is a solicited review they are willing to give this manuscript a passing grade, but strongly recommends that the matters detailed below be addressed.

As the figures are bound to be the most influential parts of this paper, let’s deal with them first.

Fig. 1. It is difficult to believe that the Avalon-White Sea transition, which for some reason is regarded as a mass origination rather than a mass extinction, represents the evolutionary appearance of nine major clades, but that’s what this diagram implies. If the vertical bars were known stratigraphic ranges, the figure would be very different. To take an extreme example, the Pentaradialomorpha is known from only one undated site in South Australia. More abundant groups, such as the Erniettomorpha, may have the stratigraphic range shown in Fig. 1, but the range extension into the Cambrian (lines 535-537) is based on two inadequately documented reports, one of which may be Ediacaran not Cambrian and the other not an erniettomorph. Furthermore, it is hard to believe that the origin of the Erniettomorpha would coincide exactly in time with the origin of the Bilateria, whether that is based on body fossils or trace fossils, but that’s what a literal reading of this figure requires. Furthermore, the oldest known erniettomorph (Pteridinium) occurs beneath a White Sea ash dated at 552.85±0.77 Ma so extending the range of the group to 560 Ma is a stretch, but it should not be presented as the exact time of origin of the whole clade. If these kinds of uncertainties were incorporated into this figure, it would be a better representation of the state of the art. Perhaps this figure should be totally recast using, for example, 95% confidence intervals on stratigraphic ranges?

Fig. 3 and the text that goes with it (lines 374-388) are merely a distraction. The negative carbon isotope excursion at ca. 500 Ma may exist in China but its meager expressions in Namibia and California-Nevada need to be confirmed before they can correlated with the poorly dated CIMP event, let alone form the basis for an extinction hypothesis. Suggest that these items be removed.

They are also confusing in that the ‘catastrophe’ that serves as a Deus ex machina for the BACE excursion/end Ediacaran extinction in Fig. 4 is attributed (unrealistically) to the Witchita igneous province and some approximately coeval volcanics in the American southwest (lines 421-425), not to the CIMP volcanism.

Fig. 4 has been redrawn to avoid conflict with the range chart of Fig. 1, but the interested reader will try to make the comparisons anyway. It seems that the alternate light and dark blue clades correspond, from left to right, to the Rangeomorpha, stem to the Porifera and/or Cnidaria; Erniettomorpha, stem to an unknown group: Dickinsoniomorpha, stem to another unknown group; and Bilateriomorpha, stem to the Bilateria (including trace and tube fossils). This is a novel reading of the record but it may be correct. However, there is no evidence from the data presented in Fig. 1 for a substantial bottleneck in the bilaterian clade, unless it is the wine glass-shaped depiction of the history of the ”Tubefauna”. The equivalent notches in the other three crown clades have even less support, since the crown group members of the Erniettomorpha and Dickinsoniamorpha have not been identified. The ‘catastrophe’ is largely artistic license.

The other problem with this figure is that it promotes the view that stem groups are doomed to fail by their very nature. This is how many paleontologists viewed the pre-extinction history of the dinosaurs prior to Alvarez et al. They were already in decline because they knew that the end of the Mesozoic was coming. More realistically, stem and crown lineages are identical at the time they split and it is only which one that survives that defines both. It is better to represent stem lineages as many equal branches of a generally diversifying tree with random odds of making it to the next level. Thus, Ediacaran lineages of crown groups are no different in principle from their sister stem lineages. Each should be given an equal probability for survival unless there is evidence to the contrary.

Lines 120-124. Seilacher’s reason for postulating a mass extinction is well explained but does not apply now since the vendobionts are not outside the Metazoa, according to recent studies. Therefore, it is misleading to say “The case for extinction was strengthened . . .”.

Lines 233+. It is good to see the addition of a section that addresses the null hypothesis, that the extinction of Ediacaran organisms was due to natural attrition not some exotic process. However, the percentage argument for mass extinction is weak, when compared with data from the Big Five during the Phanerozoic. Here, we are considering a trivial number of genera that went extinct when compared with the massive turnovers at the P-T and K-Pg boundaries. A better approach might be to ask how close to the presumed extinction horizon do the taxa in question occur? The unexpected persistence of erniettomorphs to within meters of the Ediacaran-Cambrian boundary in Namibia and Nevada provides support for the mass extinction hypothesis, but we are speaking of only 2-3 taxa, at best. Hence, this reviewer remains unconvinced. The fact that the Nama-aged Shibantan Member (lines 492-494) preserves Avalonian and White Sea taxa only reinforces this scepticism.

Review: Causes and consequences of end-Ediacaran extinction: An update — R1/PR9

Conflict of interest statement

I have worked with two of the co-authors.

Comments

The authors have addressed all my comments and concerns, and in my opinion, done a reasonable job of addressing the specific comments and concerns of the other reviewers. The result is an improved manuscript that I think can be accepted and will be a valuable contribution to ongoing investigation of the Ediacaran-Cambrian boundary.

Recommendation: Causes and consequences of end-Ediacaran extinction: An update — R1/PR10

Comments

Dear Authors:

Thank you for submitting your revised manuscript and for providing a very detailed explanation of your revisions. Your revised manuscript has been re-reviewed by two of the original reviewers (including the most critical reviewer). The more critical reviewer argues that some revisions are still required and I agree that considering those comments carefully can further improve your manuscript.

Please revise the manuscript to address the suggestions of the reviewer and provide a detailed explanation of your revisions.

I am looking forward to receiving your revised manuscript,

With best regards,

Michal Kowalewski (Handling Editor)

Decision: Causes and consequences of end-Ediacaran extinction: An update — R1/PR11

Comments

No accompanying comment.

Author comment: Causes and consequences of end-Ediacaran extinction: An update — R2/PR12

Comments

Cover Letter is attached as a ‘title page’ file, and a full response document is attached as a ‘supplementary materials’ file.

Recommendation: Causes and consequences of end-Ediacaran extinction: An update — R2/PR13

Comments

Dear Authors,

Thank you for your detailed revisions and detailed responses to comments of the reviewers and the editor. In my opinion all issues have been addressed satisfactorily and I recommend that the paper be accepted for publication in Extinction.

HE

Decision: Causes and consequences of end-Ediacaran extinction: An update — R2/PR14

Comments

No accompanying comment.